text
stringlengths
1
2.55M
id
stringlengths
21
25
metadata
dict
\section{Introduction} The relation between the quantum and classical dynamics of nonlinear systems includes a specific side in the correspondence between the dynamical properties of systems treated in the mixed and fully quantized descriptions. Various aspects of the correspondences between classical nonlinear systems on the one side and their fully quantized counterparts on the other have been intensively investigated in the last decade (see e.\ g.\ \cite{Hel91,BTU,Rei}). In many systems relevant for molecular and condensed matter physics the direct quantization of the full system in one step is, however, not possible from a practical point of view. As a rule such systems divide naturally into interacting subsystems. Then a stepwise quantization is applied resulting in a mixed description, in which one of the subsystems is treated in the quantum and the other in the classical context. Furthermore in complex systems the mixed description is often necessary for understanding global dynamical properties, e.\ g. the presence of bifurcations and separatrix structures dividing the solution manifold into characteristic parts, before for a selected energy region the full quantization can be performed. This stepwise quantization is the basic idea on which the Born - Oppenheimer approximation developed in the early days of quantum mechanics for the quantization of systems dividing into subsystems is based. As is well known this approximation can be complemented into a rigorous scheme, if the nonadiabatic couplings are included \cite{BH}. These couplings can be the source of nonintegrability and chaos of systems treated in the mixed quantum-classical description \cite{BK,BE}. Then the problem of the quantum-classical correspondences arises on the level of the relation between the dynamical properties of the mixed and fully quantized descriptions \cite{BE}. In this paper we consider this dynamic relation for the particular model of a quasiparticle moving between two sites and coupled to oscillators. This is an important model system with applications such as excitons moving in molecular aggregates and coupled to vibrations, see e.~g.\ \cite{W}. It has also attracted widespread attention in the context of the spin-boson Hamiltonian and its classical-quantum phase space behavior and correspondence (see e. g. \cite{GH,K,Sto} and references therein). Hence it seems appropriate to use this system as a model to analyze the relation between the mixed and fully quantized descriptions. Treating the oscillators in the classical or quantum contexts, whereas the quasiparticle moving between two sites is a quantum object from the beginning, one arrives at mixed and fully quantized levels of description. We have recently investigated the dynamical properties of this model in the mixed description by integrating the corresponding Bloch-oscillator equations and demonstrated the presence of a phase space with an underlying separatrix structure for overcritical coupling and chaos developing from the region of the hyperbolic point at the center of this structure. For increasing total energy chaos spreads over the product phase space of the system constituted by the Bloch sphere and oscillator plane, leaving only regular islands in the region of the antibonding states \cite{ES}. Here we consider the problem of the relation between the dynamics in the mixed and fully quantum levels of description of the coupled quasiparticle-oscillator motion. Investigating this relation we focus on the adiabatic parameter range, where the closest correspondence between the classical and quantum aspects of the oscillator dynamics can be expected. Although several aspects of the dynamics of the system have been considered \cite{GH,Sto}, there exists no systematic investigation in the adiabatic parameter range. In particular, such an investigation requires the numerical determination of a large number of eigenstates for the fully quantized system. The stationary properties of these states were reported in \cite{inc}. In this paper these states are used to compute the dynamics of the fully quantized system and to compare the quantum evolution with the dynamics of the mixed description where the oscillator is treated classically. Performing this comparison we use both the fixed and adiabatic basis sets in the mixed description. The latter basis set is of particular importance to clarify the role of the nonadiabatic couplings in the formation of the dynamics. We demonstrate the effect of the separatrix structure of the mixed description in the oscillator wave packet propagation of the fully quantized version, of dynamical subsystem correlations deriving from the separatrix structure and how the chaotic phase space regions of the system in the mixed description show up in the nonstationary properties of the time dependent full quantum state vector. In section \ref{model} the model will be specified in detail. The mixed quantum-classical description is discussed in section \ref{mqcd} including the derivation of the equations of motion, the fixed point structure and the dynamical properties of the system on this level of description. In section \ref{qevol} the evolution of the fully quantized system is presented and compared to the dynamics in the mixed description. \section{The Model}\label{model} We consider a quasiparticle coupled to oscillator degrees of freedom. The quasiparticle is specified as a molecular exciton in a tight binding representation and can be substituted by any other quantum object moving between discrete sites and described by a tight binding Hamiltonian of the same structure. The system has the Hamiltonian \begin{equation}\label{htot} H^{{\rm (tot)}} = H^{{\rm (exc)}} + H^{{\rm (vib)}} + H^{{\rm (int)}}, \end{equation} where $H^{{\rm (exc)}}$, $H^{{\rm (vib)}}$ and $H^{{\rm (int)}}$ are the excitonic, vibronic and interaction parts, respectively. $H^{{\rm (exc)}}$ represents the quantum subsystem, which is taken in the site representation \begin{equation}\label{hex} H^{{\rm (exc)}} = \sum_n \epsilon_n |c_n|^2 + \sum_{n \ne m} V_{nm} c_n^*\,c_m, \end{equation} where $c_n$ is the quantum probability amplitude of the exciton to occupy the $n$-th molecule and $V_{nm}$ the transfer matrix element. For the intramolecular vibrations coupling to the exciton we use the harmonic approximation in $H^{{\rm (vib)}}$ \begin{equation}\label{hvi} H^{{\rm (vib)}} = {1 \over 2} \sum_n (p_n^2 + \omega_n^2 q_n^2). \end{equation} Here $q_n$, $p_n$ and $\omega_n$ are the coordinate, the canonic conjugate momentum and frequency of the intramolecular vibration of the $n$-th molecule, respectively. The interaction Hamiltonian $H^{{\rm (int)}}$ represents the dependence of the exciton energy on the intramolecular configuration for which we use the first order expansion in $q_n$ \begin{equation}\label{int} H^{{\rm (int)}} = \sum_{n} \gamma_n q_n |c_n|^2, \end{equation} where $\gamma_n$ are the coupling constants. The interaction is restricted to a single oscillator at each molecule. The case of a symmetric two site system $n=1,2$, e.~g.\ an exciton in a molecular dimer constituted by two identical monomers, is considered in what follows. For this case we set $\epsilon_1=\epsilon_2$, $\omega_1=\omega_2$, $\gamma_1=\gamma_2$ and $V_{12}=V_{21}=-V$, $V>0$. Then by introducing for the vibronic subsystem the new coordinates and momenta \begin{equation}\label{qp} q_\pm := {q_2 \pm q_1 \over \sqrt{2} }\hspace*{1cm} p_\pm := {p_2 \pm p_1 \over \sqrt{2} }, \end{equation} and for the excitonic subsystem the Bloch variables \begin{equation}\label{bl} x=\rho_{21}+\rho_{12},\hspace*{1cm} y=i(\rho_{21}-\rho_{12}),\hspace*{1cm} z=\rho_{22}-\rho_{11}, \end{equation} where $\rho_{mn}$ is the density matrix of the excitonic subsystem \begin{equation}\label{dm} \rho_{mn}=c_n^*c_m, \end{equation} the relevant part of (\ref{htot})-(\ref{int}) connected with the exciton coupled to the $q_-$ vibration is obtained in the form \begin{equation}\label{hsymx} H_- = -Vx + {1 \over 2} (p_-^2 + \omega^2 q_-^2) + {\gamma q_- z \over \sqrt{2}}\,. \end{equation} The part corresponding to $q_+$ is not coupled to the exciton and omitted. The Hamiltonian (\ref{hsymx}) can be represented as an operator in the space of the two dimensional vectors $C=(c_1,c_2)$ constituted by the excitonic amplitudes $c_n$ by using the standard Pauli spin matrices $\sigma_i$ $(i=x,y,z)$. Passing in (\ref{hsymx}) to dimensionless variables by measuring $H$ in units of $2V$ and replacing $q_-$, $p_-$ by \begin{equation}\label{gqp} Q := \sqrt{2V} q_- \hspace*{1cm} P := {1 \over \sqrt{2V}} p_-, \end{equation} one finally obtains \begin{equation}\label{hsym} H = -{\sigma_x\over2} + {1 \over 2} (P^2 + r^2 Q^2) + \sqrt{p\over 2}r\,Q\sigma_z\, . \end{equation} Here \begin{equation}\label{par} p = {\gamma^2\over 2V\omega^2} \end{equation} represents the dimensionless excitonic-vibronic coupling and \begin{equation}\label{rad} r = {\omega \over 2V} \end{equation} is the adiabatic parameter measuring the relative strength of quantum effects in both subsystems. We focus on the adiabatic case $r\ll 1$ when the vibronic subsystem can be described in the classical approximation. To make contact with the dynamical features following from the adiabatic approximation we derive the basic equations in both the fixed and the adiabatic bases. \section{Mixed Quantum-Classical Description}\label{mqcd} \subsection{Fixed Basis} In this case the basis states are given by the fixed molecule sites $|n\rangle$ . Representing the excitonic state by $|\psi\rangle = \sum_n c_n |n\rangle$, inserting it into the time dependent Schroedinger equation for (\ref{hsym}) and using (\ref{bl}) the quantum equations of motion for the excitonic subsystem describing the transfer dynamics between the two sites are obtained. The classical equations for the dynamics of the oscillator are found by passing to the expectation values of $Q$ and $P$ and using (\ref{hsym}) as a classical Hamiltonian function from which the canonical equations are derived. In this way one obtains the coupled Bloch-oscillator equations representing the dynamics of the system in the mixed description \begin{eqnarray}\label{eommix} \dot{x} &=& - \sqrt{2p}\,r\,Q\, y \nonumber \\ \dot{y} &=& \sqrt{2p}\,r\,Q\, x + z \nonumber \\ \dot{z} &=& -y \\ \dot{Q} &=& P \nonumber \\ \dot{P} &=& -\,r^2 Q - \sqrt{p\over2}\,r z \nonumber \end{eqnarray} Besides the energy \begin{equation} E = -{x\over2} + {1\over2}(P^2 + r^2Q^2) + \sqrt{p\over2}rQz \end{equation} there is a second integral of the motion restricting the flow associated with the quantum subsystem to the surface of the unit radius Bloch sphere \begin{equation}\label{bln} R^2 = x^2 + y^2 + z^2 = 1\,. \end{equation} Sometimes it is advantageous to make use of this conserved quantity in order to reduce the number of variables to four, e.\ g.\ when a formulation in canonically conjugate variables is desired also for the excitonic subsystem. One then introduces an angle $\phi$ by \begin{equation}\label{defphi} x = \sqrt{1-z^2}\cos\phi\,,\hspace*{1cm}y=\sqrt{1-z^2}\sin\phi\,. \end{equation} We shall replace the usual Bloch variables by these coordinates where it is appropriate. \subsection{ Adiabatic Basis} In this case one first solves the eigenvalue problem of the part of the Hamiltonian (\ref{hsym}) which contains excitonic operators \begin{equation}\label{adham} h^{{\rm (ad)}} = -{\sigma_x\over2} + \sqrt{p\over2}rQ\sigma_z \,, \end{equation} where $Q$ is considered as an adiabatic variable. The eigenvalues of (\ref{adham}) are given by \begin{equation} \epsilon_\pm^{{\rm (ad)}}(Q)\, = \pm {1\over2}w(Q), \end{equation} where \begin{equation}\label{wu} w(Q)\, = \sqrt{1 + 2pr^2 Q^2}. \end{equation} The eigenvalues are part of the adiabatic potentials for the slow subsystem \begin{equation}\label{adpot} U_\pm^{{\rm (ad)}}(Q) = {1\over 2} r^2Q^2 + \epsilon_\pm^{{\rm (ad)}}(Q) \, , \end{equation} The two eigenstates $(\alpha=1,2)$ of (\ref{adham}) represented by the fixed basis are given by \begin{equation}\label{adbv+} |\alpha=2,Q\rangle = {1\over\sqrt{2}}(-\sqrt{1 + c(Q)}\,|2\rangle+\sqrt{1 - c(Q)}\,|1\rangle) \end{equation} and \begin{equation}\label{adbv-} | \alpha=1,Q \rangle = {1\over\sqrt{2}}(+\sqrt{1 - c(Q)}\,|2\rangle+\sqrt{1 + c(Q)}\,|1\rangle) \end{equation} with \begin{equation} c(Q) := {\sqrt{2p}rQ\over w(Q)}\,. \end{equation} The state vector of the excitonic subsystem is expanded in the adiabatic basis $|\psi\rangle = \sum_{\alpha}c^{{\rm (ad)}}_\alpha |\alpha,Q \rangle $ and inserted into the time dependent Schr\"odinger equation. For obtaining the complete evolution equations in the adiabatic basis one has to take into account the time derivative of the expansion coefficients $c^{{\rm (ad)}}_\alpha$ as well as the nonadiabatic couplings due to the time dependence of the states $|\alpha,Q(t)\rangle$. The neglect of these couplings would result in the adiabatic approximation. Using $(d/d t) |\alpha,Q \rangle = \dot Q (d/d Q)|\alpha,Q\rangle$ the nonadiabatic coupling function \begin{equation}\label{ncg} \varphi_{\alpha\beta} := \left\langle\alpha,Q\left|{\partial\over\partial Q}\right|\beta,Q \right\rangle, \end{equation} $(\varphi_{\alpha\beta}=-\varphi_{\beta\alpha})$ is found, which in case of the eigenstates (\ref{adbv+},\ref{adbv-}) is explicitly given by \begin{equation}\label{nce} \varphi_{12} = - \frac{\sqrt{p}r}{\sqrt{2}[w(Q)]^2}\,. \end{equation} Introducing now analogously to (\ref{bl}) the Bloch variables in the adiabatic basis and treating the oscillator in the classical approximation one obtains the coupled Bloch-oscillator equations in the adiabatic basis \begin{eqnarray}\label{eomad} \dot x^{{\rm (ad)}} &=& 2P\, \varphi_{12}(Q)\, z^{{\rm (ad)}} - w(Q)\, y^{{\rm (ad)}}\nonumber \\ \dot y^{{\rm (ad)}} &=& w(Q)\, x^{{\rm (ad)}}\nonumber \\ \dot z^{{\rm (ad)}} &=& -2P\, \varphi_{12}(Q)\, x^{{\rm (ad)}}\\ \dot Q &=& P\nonumber \\ \dot P &=& -r^2Q + \sqrt{p}\,w(Q)\varphi_{12}\,x^{{\rm (ad)}} - \sqrt{p\over2}\,r\,c(Q)\,z^{{\rm (ad)}}\nonumber \end{eqnarray} The energy is now expressed by the adiabatic Bloch variable $z^{{\rm (ad)}}$ \begin{equation}\label{ad1dz} E = w(Q){z^{{\rm (ad)}}\over2} + {1\over 2}\left(P^2 + r^2Q^2\right) \end{equation} and the flow is again located on the surface of the unit Bloch sphere. Neglecting the nonadiabatic couplings $\varphi_{12}=0$ one obtains the dynamics of the decoupled adiabatic oscillators. The adiabatic oscillators can be considered as one dimensional integrable subsystems corresponding to the Hamiltonians \begin{equation}\label{ad1d} h_\pm^{{\rm (ad)}} = {1\over 2} P^2 + U_\pm^{{\rm (ad)}}(Q) \, , \end{equation} where $U_\pm^{{\rm (ad)}}(Q)$ is given by (\ref{adpot}). The connection between the Bloch variables in the fixed and the adiabatic basis is given by\\[3mm] \begin{tabular}{ll} $ x = - c(Q) x^{{\rm (ad)}} - \sqrt{1-c(Q)^2}\,z^{{\rm (ad)}} \hspace*{1cm}$& $ x^{{\rm (ad)}} = - c(Q) x - \sqrt{1-c(Q)^2}\,z $\\ $ y = -y^{{\rm (ad)}} $& $ y^{{\rm (ad)}} = -y $\\ $ z = c(Q) z^{{\rm (ad)}} - \sqrt{1 - c(Q)^2} x^{{\rm (ad)}} $& $ z^{{\rm (ad)}} = c(Q) z - \sqrt{1 - c(Q)^2} x $ \end{tabular} \begin{equation}\label{trbloch} \sqrt{1-z^2}\,\sin\phi=-\sqrt{1-(z^{{\rm (ad)}})^2}\,\sin\phi^{{\rm (ad)}} \end{equation} Using these transformation formulas one can show that the equations of motion (\ref{eommix}) derived in the fixed basis are actually equivalent to those in the adiabatic basis (\ref{eomad}). \subsection{Fixed points and bifurcation} Essential information about the phase space of the excitonic-vibronic coupled dimer is contained in the location and the stability properties of the fixed points of the mixed quantum-classical dynamics. Setting in the equations of motion for the fixed basis (\ref{eommix}) all time derivatives to zero, we find for any stationary state $$ Q_s = -{1\over r}\sqrt{p\over2}z_s \,,\hspace*{1cm} P_s = 0 \,,\hspace*{1cm} y_s = 0 $$ \begin{equation}\label{statcon} z_s - p x_s z_s = 0 \,. \end{equation} The stability properties of a fixed point are determined by a linearization of the equations of motion using canonical variables \cite{ES}. It is appropriate to subdivide all stationary points according to whether they are located in the bonding region $x_s > 0$ or in the antibonding region $x_s < 0$. There is no transition between these two groups when the parameters of the system are varied since $x_s = 0$ is excluded by (\ref{statcon}). This terminology is in accordance with molecular physics where it is common to refer to the state $x=1$ with symmetric site occupation amplitudes $c_1 = c_2$ as bonding and to the state $x=-1$ with antisymmetric amplitudes $c_1=-c_2$ as antibonding. \subsubsection{Bonding region ($x_s > 0$)} We consider the bonding region first. The location of the fixed points is obtained from (\ref{statcon}) using the additional restriction \begin{equation} x_s^2 + z_s^2 = 1\,. \end{equation} One finds the following solutions in dependence on the value of the dimensionless coupling strength $p$: \begin{description} \item{(A) $0 \le p \le 1$:} In this case (\ref{statcon}) allows for a single solution only. \begin{equation} {\bf g:}\hspace*{1cm} x_s = 1 \,,\hspace*{1cm} z_s = 0 \,,\hspace*{1cm} Q_s = 0 \,,\hspace*{1cm} E_s =-{1\over 2} \end{equation} This point is the bonding ground state corresponding to a symmetric combination of the excitonic amplitudes $c_1 = c_2 = 1/\sqrt{2}$. $\bf g$ is stable elliptic. \item{(B) $p \ge 1$:} A bifurcation has occurred and we obtain three stationary points. \begin{equation} \begin{array}{ccc} {\bf g_{\pm}:} & x_s = {1\over p} & z_s = \pm{\sqrt{p^2-1}\over p} \\ & Q_s=\pm{\sqrt{p^2-1}\over\sqrt{2p\,}r} & E_s= -{p^2+1\over4p} \end{array} \end{equation} These two points are stable elliptic. \begin{equation} {\bf h:}\hspace*{1cm} x_s = 1\,,\hspace*{1cm} z_s =0 \,,\hspace*{1cm} Q_s = 0 \,,\hspace*{1cm} E_s = -{1\over 2} \end{equation} The point $\bf h$ is at the position of the former ground state, but in contrast to $\bf g$ it is unstable hyperbolic. \end{description} \noindent The parameter $p$ governs a pitchfork bifurcation: The ground state g below the bifurcation ($p < 1$) splits into two degenerate ground states $\bf g_\pm$ above bifurcation ($p > 1$). At the former ground state a hyperbolic point $\bf h$ appears. This situation is also obvious from fig.~\ref{adpotplot}(a). \subsubsection{Antibonding region ($x_s < 0$)} Independent on the coupling strength $p$ we have in this region only one solution of (\ref{statcon}) (see fig.\ \ref{adpotplot}(b)): \begin{equation} {\bf e:} z_s = 0 \,,\hspace*{1cm} x_s= -1 \,,\hspace*{1cm} Q_s = 0 \,,\hspace*{1cm} E_s = +{1\over 2} \end{equation} This stationary state corresponds to an antisymmetric combination of the excitonic amplitudes $c_1 = -c_2 = 1/\sqrt{2}$. $\bf e$ is stable for \begin{equation}\label{stababond} \frac{|r^2 - 1|}{r} > 2\sqrt{p}\,, \end{equation} which holds when the system is not in resonance and in particular for the adiabatic case $r \ll 1$. \paragraph*{} Since the equations of motion in the fixed and in the adiabatic basis are equivalent, it is clear that the fixed same points (\ref{statcon}) can also be obtained from (\ref{eomad}). Setting in (\ref{eomad}) the time derivatives of $x$, $y$ and $Q$ equal to zero, one finds for the stationary states of the adiabatic case \begin{eqnarray} \label{ans} x_s^{{\rm (ad)}} = 0 \,,\hspace*{1cm} y_s^{{\rm (ad)}} = 0 \,,\hspace*{1cm} P_s^{{\rm (ad)}} = 0\,, \end{eqnarray} leaving for the stationary values of $z_s^{{\rm (ad)}}$ the poles \begin{equation}\label{adz} z_s^{{\rm (ad)}} = \pm 1 \end{equation} It is worth noting that in the adiabatic basis the stationary states are always located at $z^{{\rm (ad)}}_s=\pm 1$ and this will be the case for any system treated in mixed quantum-classical description and restricted to the two lowest adiabatic levels. A specific feature of using the adiabatic basis is the independence of the location of the fixed points on the explicit form of the nonadiabatic coupling function since this function enters the equations of motion in form of the products $\varphi_{12}(Q)P$ and $\varphi_{12}(Q)x^{{\rm (ad)}}$ which drop out at a fixed point because of (\ref{ans}) and (\ref{adz}). From (\ref{trbloch}) it is moreover easy to see that the fixed points in the bonding region are located within the lower adiabatic potential while the antibonding fixed points belong to the upper one. The eq.\ $\dot {P}=0$ reduces to \begin{equation}\label{adQ} (w(Q) + pz_s^{{\rm (ad)}})Q=0\,. \end{equation} For $z^{{\rm (ad)}}_s=+ 1$ the only solution of (\ref{adQ}) is $Q_s=0$, whereas for $z^{{\rm (ad)}}_s=-1$ one obtains additional solutions for $p>1$. These solutions are easily seen to correspond to the bifurcation discussed above. \subsection{Integrable Approximations} Before we investigate the dynamics of the complete coupled equations of motion (\ref{eommix}) or (\ref{eomad}) we would like to mention two integrable approximations to the model. The first and trivial integrable approximation is to set in the equations of motion in the fixed basis (\ref{eommix}) $p=0$ which results in a decoupling of the excitonic and vibronic motions. The second and more interesting integrable approximation is obtained by neglecting the nonadiabatic coupling function $\varphi_{12}=0$ in the equations of motion (\ref{eomad}) of the adiabatic basis which defines the adiabatic approximation from a dynamical point of view. In this approximation some of the nonlinear features of the model are still contained in the integrable adiabatic reference oscillators (\ref{ad1d}). In particular the lower adiabatic potential (\ref{adpot}) displays the bifurcation from a single minimum structure to the characteristic double well structure when the parameter $p$ (\ref{par}) passes through the bifurcation value $p=1$. It is also important to note that the fixed point structure is not changed when the nonadiabatic couplings are switched off: Neglecting $\varphi_{12}$ in (\ref{eomad}) results in the same fixed points equations as in the case including the nonadiabatic couplings. From the formal side the neglect of $\varphi_{12}$ not necessarily leads to $z_s^{{\rm (ad)}} = \pm 1$: According to the equations of motion(\ref{eomad}) $\varphi_{12}=0$ implies $z_s^{{\rm (ad)}} = const$. Then in the dynamics of the adiabatic approximation both adiabatic modes can be occupied and only the transitions between them are switched off. The oscillator equations become autonomous describing regular motions according to the classical Hamilton function (\ref{ad1dz}) with $z_s^{{\rm (ad)}}$ as a parameter. The oscillator coordinate $Q(t)$ enters the Bloch equations for $x(t)$ and $y(t)$. The equations for the latter describe the regular motion on a circle generated by an intersection of the Bloch sphere with the plane $z_s^{{\rm (ad)}} = const$ on which the phase oscillations between the modes are realized. In the following we demonstrate that the regular structures associated with the adiabatic approximation are present in both the mixed and fully quantized descriptions. At the same time we show that the complete coupled system of Bloch-oscillator equations, i.~e.\ including the nonadiabatic couplings, displays dynamical chaos. This identifies the nonadiabatic couplings as a source of nonintegrability and chaos in the mixed description of the system and rises the question about the signatures of this chaos after full quantization is performed. The latter problem will be addressed in the last section. \subsection{Dynamical Properties} \label{sec:dynmix} The dynamical properties of the coupled Bloch-oscillator equations (\ref{eommix}) were analyzed by a direct numerical integration. Some of our results, such as the presence of chaos in the mixed description of the excitonic-vibronic coupled dimer, were reported in \cite{ES}. Therefore the aim of this section is twofold: On the one side we reconsider the findings in \cite{ES} relating the dynamical structures to the adiabatic approximation, in which the integrable reference systems (\ref{ad1d}) can be defined. This clarifies the role of the nonadiabatic couplings in the formation of the dynamics of the model, which was not done before. On the other side we provide the necessary characterization of the phase space structure, such as the location of the separatrix dividing the phase space into trapped and detrapped solutions, and the identification of the regions and associated parameters belonging to the regular and chaotic parts of the dynamics, respectively. The latter points will provide the basis to perform the comparison of the mixed description with the full quantum evolution in the next section. The numerical integration of the equations of motion in the mixed description can be performed both in the fixed (\ref{eommix}) and the adiabatic basis (\ref{eomad}). The integration in the fixed basis, however, is numerically simpler and the fixed Bloch variables provide a more convenient frame for the excitonic motion. Hence we used the fixed basis for a numerical integration. It must be stressed, however, that the representation of the excitonic system in the fixed and the adiabatic basis are equivalent, if the nonadiabatic couplings are included. The connection between both representations is given by the eqs. (\ref{trbloch}). In view of the existence of two integrals of the motion three variables from the total of five variables of the system are independent. Therefore a standard two dimensional Poincar\'e surface of section is defined by fixing one variable. According to the choice of variables sections can be defined for both the oscillatory and excitonic subsystems. The dynamics of the system was found to depend crucially on the choice of the total energy with respect to the characteristic energies of the system such as the minima of the adiabatic potentials and above the bifurcation the energy corresponding to the hyperbolic fixed point $E_h$. Above the bifurcation the separatrix structure, which divides the phase space into characteristic parts, is present. The regular phase space structure following from the integrable adiabatic reference oscillators (\ref{ad1d}) above the bifurcation is shown in fig.\ \ref{adpotplot}. In fig.\ \ref{pq-0.8}(a) a Poincar\'e section in oscillator variables is presented for the value $p=0.8$ which is below the bifurcation value $p=1$. In this case the adiabatic potential $U_-(Q)$ has a single minimum. The total energy is chosen at $E=0$, i.\~e.\ well below the minimum of the upper adiabatic potential. Therefore the influence of this potential is small and the oscillator dynamics can be expected to be close to the regular dynamics of the lower reference oscillator associated with $U_-(Q)$. This is indeed confirmed by fig.\ \ref{pq-0.8}(a). There is, however, a chain of small resonance islands in the outer part of the section, which is due to resonance between the oscillator motion and the occupation oscillations between the adiabatic modes corresponding to the finite $\dot z^{{\rm (ad)}}$ in the case of the presence of the nonadiabatic couplings. The interaction between the occupation oscillations due to the finite $\dot z^{{\rm (ad)}} $ and the oscillator motion becomes much more pronounced for higher energies. A corresponding Poincar\'e section is displayed in fig.\ \ref{pq-0.8}(b) for the same value $p=0.8$ of the coupling constant, but with the energy now chosen above the minimum of the upper adiabatic potential. This choice of the energy allows according to (\ref{ad1dz}) for a much broader range for the variation of the variable $z^{{\rm (ad)}}$ and consequently the nonadiabatic couplings are more effective. Correspondingly we observe now several resonance chains. Increasing the coupling above the bifurcation value $p=1$, but fixing the total energy below $E_h$ one expects regular oscillations around the displaced minima of the double well structure in $U_-(Q)$. A Poincar\'e section in the oscillator variables corresponding to this behavior is shown in fig.\ \ref{pq-3.4}(a), where $p=3.4$ and the total energy is below the saddle point of the potential $U_-(Q)$. Increasing the energy to a value slightly above $E_h$ one finds sections displaying oscillations resembling the separatrix structure as shown in fig.\ \ref{pq-3.4}(b). For energies well above $E_h$ chaotic trajectories do exist. Characteristic examples are provided by the Poincar\'e sections in oscillator variables displayed in fig.\ \ref{pq-3.4}(c) and (d), where the regions of regular and chaotic behavior of the oscillator subsystem are shown for two cases of total energy above $E_h$. In the case (c) the total energy is below the case (d). It is seen how with increasing energy the regular part of the oscillator phase space becomes smaller and the chaotic part increases. The corresponding regular and chaotic components of the excitonic subsystem are located in the antibonding and bonding regions of the Bloch sphere, respectively (see also fig.\ \ref{03} below). Relating the location of the dynamics on the Bloch sphere to the energy of the excitonic subsystem we find that for chaotic trajectories the excitonic subsystem is in an energetically low state within the bonding region of the Bloch sphere whereas for regular trajectories the excitonic subsystem is in its energetically high state within the antibonding region. Correspondingly, the energy of the vibronic subsystem is high for chaotic dynamics and low for regular dynamics, because the total energy is the same for all the trajectories displayed in each of the figures \ref{pq-3.4}(c) and (d). Hence the destruction of the regular dynamics is connected with the energy of the vibronic subsystem: Regular dynamics is realized for oscillator states with low energy, small amplitude oscillations and consequently small effective coupling, whereas high oscillator energy destroys the regular structures and results in global chaos. The dynamics of the oscillator subsystem is complemented by the Poincar\'e sections on the surface of the Bloch sphere showing the behavior of the excitonic subsystem. In the figs.\ \ref{03}(a)-(d) such a typical set of Poincar\'e sections is presented for different energies and above the bifurcation ($p = 2.0$). The sections correspond to the left turning point of the oscillator. For low energy one finds regular dynamics in the region of the bifurcated ground states. These regular trajectories represent the self trapped solutions of the system in which the exciton is preferentially at one of the sites of the dimer and correspond to the one sided oscillations of the vibronic subsystem of fig.\ \ref{pq-3.4}(a). Increasing the energy local chaos starts in the vicinity of the hyperbolic point $E_h$. The local chaos can be considered as a perturbation of the dynamics near the saddle of the lower potential $U_-(Q)$ due the nonadiabatic couplings of the adiabatic oscillators. With increasing energy chaos spreads over the Bloch sphere leaving only regular islands in the region of antibonding states associated with the upper adiabatic potential and in accordance with the dynamics of the vibronic subsystem discussed above. For high enough energy the coupling between the adiabatic reference oscillators almost completely destroys regular structures and results in global chaos. \section{Quantum Evolution}\label{qevol} We now turn to the dynamics in the full quantum description of the model considering in the Hamiltonian (\ref{hsym}) the coordinate $Q$ and the momentum $P$ as non-commuting quantum variables. We focus on the features of the evolution in the adiabatic parameter region for $r\ll 1$. The evolution of the full quantum state vector of the system satisfying some fixed initial condition is computed from the eigenstate representation of the Hamiltonian. For a realistic description of the system in the adiabatic parameter region a large number of eigenstates had to be used in the expansion. Correspondingly, the diagonalization of the Hamiltonian (\ref{hsym}) with $Q$ and $P$ being quantum operators was performed using a large set of oscillator eigenfunctions for the undisplaced oscillator as a basis, i.~e.\ the basis was constructed from the product states $|n,\nu\rangle:=|n\rangle \otimes |\nu\rangle$, where the index $n=1,2$ labels the two sites of the dimer and $\nu=0,1,\dots$ stands for the oscillator quantum number. In this basis the quantized version of the Hamiltonian (\ref{hsym}) is represented by the matrix \begin{equation} \langle n,\nu| H | n',\nu '\rangle = -{[1-(-1)^{\nu+\nu'}]\over 4}\delta_{\nu ,\nu'} \,+\, r \left( \nu + {1 \over 2}\right)\delta_{n, n'}\delta_{\nu,\nu '} \,+ \end{equation} \begin{equation}\label{matel} +\, {\sqrt{p\,r}\over 2}(-1)^n\,(\sqrt{\nu'}\delta_{\nu,\nu ' -1} + \sqrt{\nu}\delta_{\nu ,\nu ' + 1})\,\delta_{n, n'}\,. \end{equation} The typical number of oscillator eigenfunctions used was $750$ yielding a total of $1500$ basis states. The properties of the stationary eigenstates, the fine structure of the spectrum and in particular the influence of the adiabatic reference oscillators and the role of the nonadiabatic couplings in the formation of the spectrum were reported in \cite{inc}. Here we consider the nonstationary properties of the full quantum system based on this eigenstate expansion and demonstrate how the nonlinear features of the dynamics in the mixed quantum-classical description are reflected in the time dependence of the full quantum state vector $|\Psi(t)\rangle$. We investigated the evolution of wave packets initially prepared in the product state \begin{equation}\label{wp} |\Phi,\alpha\rangle = |\Phi_{z_0,\phi_0}\rangle \otimes |\alpha_{Q_0,P_0}\rangle \end{equation} where $\Phi$ is an excitonic two component wave function which is specified up to an irrelevant global phase by the expectation values of the Bloch variables $z$ and $\phi$ (see (\ref{defphi})). $\alpha$ represents a standard coherent state in the oscillator variables, which is specified by the complex parameter \begin{equation}\label{aqp} \alpha(Q,P) = \sqrt{r\over2}\langle\alpha|\hat Q|\alpha\rangle + {i\over\sqrt{2r}}\langle\alpha|\hat P|\alpha\rangle \end{equation} with $Q$ and $P$ being the corresponding expectation values of position and momentum. In order to map the motion of the full state vector $|\Psi(t)\rangle$ constructed from the eigenstate expansion according to the initial condition onto an analogue of the phase space of the mixed description, in which the oscillator is treated classically, we used for the oscillator subsystem the Husimi distribution, which is an appropriate quantum analogue to the classical phase space distribution (see e.\ g.\ \cite{Tak}). It is defined by projecting $|\Psi(t)\rangle$ on the manifold of coherent states \begin{equation}\label{husdist} h_{z,\phi}(Q,P):=|\langle\Phi_{z,\phi},\alpha_{Q,P}|\Psi(t)\rangle|^2, \end{equation} where now $Q$ and $P$ are varied in the oscillator plane while $z$ and $\phi$ are fixed parameters. Without the interaction between the subsystems a wave packet prepared in a coherent oscillator state would travel undistorted along the classical trajectory started at $(Q_0,P_0)$. A weak coupling below the bifurcation ($p<1$) results in a similar picture (not displayed) with the wave packet after some initial period almost uniformly covering the classical trajectories such as those displayed in figs.\ \ref{pq-0.8}(a) and (b). Of particular interest is the effect of the separatrix structure characterising the mixed description above the bifurcation ($p>1$) on the propagation of the oscillator wave packet. For a system with a proper classical limit and a separatrix in the classical phase space the correspondence to the quantum evolution was studied e.~g.\ in \cite{Rei}. Similar to this we found that the presence of the separatrix is clearly reflected in the wave packet dynamics when the energy is fixed at $E_h$. In the set of figs.\ \ref{wp-ini} the evolution of a quantum state prepared initially right at the hyperbolic fixed point $\bf h$ is presented. The relevant system parameters are $p=2$ and $r=0.01$ and the Husimi distribution (\ref{husdist}) for the projection onto the excitonic state $z=0$, $\phi=0$ is displayed. It is seen how the oscillator wave packet spreads along the unstable direction of the separatrix structure. The asymmetric distortion of the wave packet in the beginning of the propagation, when the support of the Husimi distribution is given by the unstable direction of the separatrix, is remarkable. For long times the wave packet covers the separatrix structure more uniformly (see fig.\ \ref{wp-lt}(a)). In the set of figs. \ref{hus-hp} contour plots for an analogous wave packet propagation started at the hyperbolic point but for a larger adiabatic parameter $r=0.1$ are presented. The propagation along the separatrix structure, which is now indicated by a full line, is again evident. This indicates that the well known classical-quantum correspondence in the case of regular dynamics, namely that the quantum distribution corresponds to the orbit of the corresponding classical system, can be extended to systems treated in a mixed quantum-classical description. A more detailed comparison of the results for $r=0.01$ (fig.\ \ref{wp-ini}) and $r=0.1$ (fig.\ \ref{hus-hp}) reveals as expected that the width of the wave packet transversal to the underlying classical structure is reduced as the system is closer to the adiabatic limit. We conclude that in the adiabatic regime regular structures such as a separatrix in the formally classical phase space of the mixed description can serve to forecast qualitatively the evolution of a wave packet in the fully quantized system. In fig.\ \ref{wp-lt} we compare the Husimi distributions for one and the same wave packet projected onto two different excitonic states in order to reveal the quantum correlations between the excitonic and the vibronic subsystems. In fig.\ \ref{wp-lt}(a) we chose $z=0$ and $\phi=0$ corresponding to equal site occupation probabilities whereas in fig.\ \ref{wp-lt}(b) the wave packet is projected onto $z=1$, i.~e.\ an excitonic state completely localized at one of the dimer sites. It is seen that for the case of an equal site occupation the oscillator evolution proceeds along both branches of the separatrix structure whereas for the one sided projection the oscillator is preferentially located on the branch of separatrix corresponding adiabatically to the occupied site. This behavior reflects the property of the quantum system to include coherently all the variants of motion of the mixed quantum-classical system weighted with the corresponding probability in analogy to the semiclassical propagator of a system with proper classical limit, which is given as a sum over classical trajectories. Finally we address the problem of how the qualitative differences between the regular and the chaotic dynamics of the system in the mixed description are reflected in the evolution of the fully quantized system, i.~e.\ whether there are signatures of the dynamic chaos in the mixed description in the time dependent state vector of the fully quantized system. For simple systems quantized in one step and chaotic in the classical limit the differences in the quantum evolution between initial conditions selected in the classical regular and chaotic parts of the phase space of the system are well known: If e.~g.\ the initial conditions of the quantum system are selected in the regular part of the classical phase space the time dependence of the appropriately chosen quantum expectation values follow ("shadow") the corresponding classical values over a substantial amount of time, whereas for initial conditions chosen in the chaotic part of the classical phase space these dependences start to deviate from each other almost immediately (see e.\ g.\ \cite{BBH}). In order to investigate this connection in our case we have selected different initial conditions in the regular and chaotic parts of the Bloch sphere of the system and compared the evolution in the mixed description with that of expectation values obtained from the fully quantized system. In fig.\ \ref{icond} the location of three different initial conditions on the Bloch sphere of the excitonic subsystem belonging to the main regular (A) and chaotic (B) regions of the dynamics in the mixed description, as well as a small regular island (C) embedded in a large chaotic surrounding are shown. For a comparison of the dynamics in the mixed and fully quantized descriptions for these cases we selected the variables $Q(t)$ and $z(t)$ displayed in the upper parts in the set of figs.\ \ref{tdepA}-\ref{tdepC}. We first compare the dynamics for initial conditions located in the main regular (antibonding) and main chaotic (bonding) regions. Since the initial state of the fully quantized system is chosen as a product state with factorizing expectation values for which the decoupling implicit in the derivation of (\ref{eommix}) is justified, there is always an interval at the beginning of the time evolution where the mixed description follows closely the quantum data. Then, however, there is indeed a striking difference between initial conditions selected in the regular and the chaotic parts of the phase space of the mixed system: For initial conditions in the regular part ({\bf A}, fig.\ \ref{tdepA}) the quantum expectation value $Q(t)$ follows closely the classical trajectory of the mixed description over several periods and then, apart from a slowly growing phase shift, both dependences keep a similar oscillatory form, whereas for an initial condition in the chaotic part ({\bf B}, fig.\ \ref{tdepB}) the corresponding curves are completely different and the deviation between both starts already after a fourth of the oscillator period. This confirms for our case the general behavior of classically chaotic systems to produce a fast breakdown of the validity of quasiclassical approximations when quantum effects become important. The comparison for the occupation difference $z(t)$ of the excitonic sites is not so direct, because the exciton constitutes the fast subsystem resulting in rapid oscillations of $z(t)$ in the mixed description. However, for the regular case we observe that the slowly changing mean value of $z(t)$ obtained from the mixed description is related to the quantum data, though amplitude and phase of the superimposed rapid oscillations are different after a few periods of the excitonic subsystem. In the chaotic case the breakdown of the mixed description for shorter times is evident and there is no correspondence for the mean values. The gradual development of quantum correlations between both subsystems, which are absent in the initially factorized state, can be quantified by calculating the effective Bloch radius \begin{equation}\label{qbr} R(t)=\sqrt{x(t)^2+y(t)^2+z(t)^2} \end{equation} of the excitonic subsystem using the time dependent expectation values of $\sigma_x$, $\sigma_y$ and $\sigma_z$. Note that the reduced density matrix $\rho(t)$ for the excitonic subsystem, obtained from the full density matrix by taking the trace over the oscillator states, is related to $R(t)$ via ${\rm Tr}\, \rho(t)^2=(1/2)(1 + R(t)^2)$. For the factorized and correspondingly uncorrelated initial quantum state the value of the Bloch radius is $R=1$ and $R(t)$ will decrease in the course of time according to the degree to which quantum correlations lead to an entanglement between both subsystems. In the lower parts of the figs.\ \ref{tdepA}-\ref{tdepC} the dependence of the Bloch radius on the time is displayed for a long time interval. The difference between the behavior for initial conditions chosen in the regular and chaotic parts of the phase space of the mixed description is remarkable: For initial conditions in the regular part of the phase space after an initial drop $R(t)$ stabilizes at a value close to $1$, whereas for the initial conditions in the chaotic part the descent is much more pronounced and the long time value of $R(t)$ is much lower, thus indicating stronger quantum correlations in the chaotic case. The correlations between the subsystems are the reason for the breakdown of the mixed description which implicitly contains the factorization of expectation values. Therefore the smaller value of $R(t)$ observed for the state prepared in the chaotic region confirms the faster breakdown of the mixed description as compared to a regular initial state. However, it is important to note that the striking difference between the values of $R(t)$ is not restricted to this initial period but extends to much longer times (which are on the other hand small compared to the time for quantum recurrences). In this respect our results indicate time dependent quantum signatures of chaos of the mixed description which are beyond the well known different time scales for the breakdown of quasiclassical approximations. Finally we present the example for a quantum state prepared on a regular island embedded into chaotic regions of the mixed quantum-classical phase space ({\bf C}, fig.\ \ref{tdepC}). The structure of the selected island is shown in the lower part of fig.\ \ref{icond}. For such a state the situation is specific due to the spreading of the quantum state out of the regular island. After some initial time in which the quantum dynamics probes the regular region of the the mixed dynamics the wave packet enters the region in which the mixed dynamics is chaotic. Correspondingly we find for an initial time interval that the agreement between the mixed and the full quantum description is as good as expected for regular dynamics whereas for long times the quantum system shows the typical behavior of a chaotic state. This is evident from the time dependence of the Bloch radius which is displayed in the lower part of fig.\ \ref{tdepC} on a sufficiently large time scale. \pagebreak \section{Conclusions} 1. We considered the nonlinear dynamical properties of a coupled quasiparticle-oscillator system and demonstrated that the regular structures of the mixed quantum-classical description such as the fixed points and the presence of a separatrix are associated with the corresponding adiabatic approximation, in which the nonadiabatic couplings are switched off and integrable reference systems can be defined. Comparing the evolution of quantum wave packets to the mixed quantum-classical description we found that regular structures of the mixed description can serve as a support for wave packet propagation in the fully quantized system in the adiabatic regime. This should be of interest for other systems to which a stepwise quantization must be applied due to their more complex structure, e.~g.\ for the purpose of forecasting the qualitative properties of propagating wave packets using the mixed description as a reference system. 2. The nonadiabatic couplings, the inclusion of which is beyond the adiabatic approximation, are identified as the source of dynamical chaos observed in the mixed quantum-classical description. This suggests that nonadiabatic couplings can be a general source of nonintegrability and chaos also in other systems treated along a stepwise quantization. Signatures of this type of chaos can then be expected on the fully quantized level of description similar to what we found for the coupled quasiparticle-oscillator system. In particular, the breakdown of the mixed description is enhanced for states prepared in a chaotic region of the phase space and the long time evolution of these states is characterized by much stronger quantum correlations between the subsystems. 3. Our results are related to the general question of how the idea of the Born-Oppenheimer approach to analyze complex systems by a stepwise quantization can be extended to a dynamical description. A more systematic investigation of this question using other model systems is certainly of interest in view of the widespread use of this approach. \section{Acknowledgement} Financial support from the Deutsche Forschungsgemeinschaft (DFG) is gratefully acknowledged. \pagebreak
proofpile-arXiv_065-400
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} In the last few years numerous theoretical and experimental studies have focused on the electron spectrum in semileptonic inclusive $B\to X_c\ell\bar{\nu_{\ell}}$ decays. The electron spectrum from free quark decays receives both perturbative and nonperturbative corrections. Knowledge of the shape of the spectrum can provide insights into nonperturbative effects in $B$ meson decays, and thereby also give some information on the weak mixing angle $|V_{cb}|$. In the framework of Heavy Quark Effective Theory (HQET) it is possible to show that the quark level decay rate is the first term in a power series expansion in the small parameter $\Lambda_{QCD}/m_b$ \cite{CGG}. For infinitely heavy quarks the free quark model is an exact description of heavy meson physics. At finite quark masses the first few terms in the heavy quark expansion have to be taken into account. Expressions for these nonperturbative corrections to the lepton spectrum are known to order $(\Lambda_{QCD}/m_b)^3$ \cite{mark,mannel,bd,gk} and the ${\cal O}(\alpha_s)$ perturbative corrections to the free quark decay were given in \cite{jk}. The dominant remaining uncertainties are the two-loop corrections to the quark level decay rate and the perturbative corrections to the coefficients of the HQET matrix elements in the operator product expansion. Here we examine the former. While a full two-loop calculation of the electron spectrum is a rather daunting task, it is possible to calculate the piece of the two-loop correction that is proportional to $\beta_0=11-2/3 n_f$ with relative ease by performing the one-loop QCD corrections with a massive gluon. The $\alpha_s^2\beta_0$ parts of the two-loop correction may then be obtained from a dispersion integral over the gluon mass\cite{sv}. If there are no gluons in the tree level graph, the $\alpha_s^2\beta_0$ part of the two-loop contribution is believed to dominate the full $\alpha_s^2$ result because $\beta_0$ is rather large. Several examples supporting this belief are listed in \cite{asmark}. A recent calculation \cite{asmark} of the $\alpha_s^2\beta_0$ correction to the total inclusive rate for $B\to X_c\ell\bar{\nu_{\ell}}$ decays showed that the $\alpha_s^2\beta_0$ parts of the two-loop correction are approximately half as big as the one-loop contribution, resulting in a rather low BLM scale \cite{BLM} of $\mu_{BLM}=0.13m_b$. For the electron spectrum we find that this part of the second order correction also amounts to about 50\% of the order $\alpha_s$ contribution, at all electron energies except those close to the endpoint. Close to the endpoint the corrections are roughly equal in magnitude. The HQET matrix elements $\lambda_1$ and $\bar\Lambda$ can be extracted from the electron spectrum in $B\to X_c\ell\bar{\nu_{\ell}}$ decays \cite{gklw}. Even though this method of obtaining HQET matrix elements was found to be rather insensitive to the first order perturbative corrections, it is useful to extract $\lambda_1, \bar\Lambda$ including the $\alpha_s^2\beta_0$ corrections. Then these matrix elements can be used to relate the pole quark mass to the $\overline{\rm MS}$ masses at order $\alpha_s^2\beta_0$. Similarly, one can include the $\alpha_s^2\beta_0$ parts of the two-loop contribution in the theoretical prediction for the total rate, which is needed for the determination of $|V_{cb}|$. Since the $\alpha_s^2\beta_0$ corrections are rather large the resulting changes in the quark masses and $|V_{cb}|$ are not negligible. In Sect.~II and III we give analytic expressions for the contributions from virtual and real gluon radiation. The last phase space integral in the virtual correction and the last two integrals in the bremsstrahlung are done numerically. Readers not interested in calculational details are advised to skip these sections. In Sect.~IV we combine the results from the previous two sections to obtain the $\alpha_s^2\beta_0$ corrections to the electron spectrum, and discuss the implications for the extraction of $\bar\Lambda,\lambda_1$, the $\overline{\rm MS}$ quark masses, and $|V_{cb}|$. In Appendix \ref{ap2} we give an interpolating polynomial which reproduces the two-loop correction calculated here. \section{Virtual Corrections} The corrections from massive virtual gluons can be calculated in complete analogy to the usual one-loop QCD corrections. The ultraviolet divergence in the vertex correction cancels when combined with the quark wave function renormalizations. There is no infrared divergence since we do the calculation with a massive gluon. The virtual one-loop correction to the differential rate can be written as \begin{eqnarray} \frac{ {\rm d} \Gamma^{(1)}_{virt}(\hat{\mu}) }{ {\rm d} y}&=&\alpha_s^{(V)} \frac{|V_{cb}|^2G_F^2 m_b^5}{48 \pi^4}\int {\rm d} \hat{q}^2 \Big[ 2(y- \hat{q}^2 )( \hat{q}^2 +1-r^2-y) (a_1+a_{wr}) -2 r \hat{q}^2 a_2 \nonumber \\ &&+( \hat{q}^2 (y-1)+y(1-r^2)-y^2)a_3\Big] \label{v1loop} \end{eqnarray} where $\hat{\mu}=\mu/m_b$ is the rescaled gluon mass, and $y=2E_e/m_b$, $r=m_c/m_b$, and $ \hat{q}^2 =q^2/m_b^2$ are the rescaled electron energy, charm mass, and momentum transfer, respectively. The limits for the integration over $ \hat{q}^2 $ are \begin{equation} 0\le \hat{q}^2 \le\frac{y(1-y-r^2)}{1-y}, \qquad 0\le y\le 1-r^2. \end{equation} The functions $a_{wr}( \hat{q}^2 )$ and $a_i( \hat{q}^2 ),\ i=1,2,3$ are the contributions from the wave function renormalization and the vertex correction respectively. They can be expressed in terms of the scalar two- and three-point functions $B_0$ and $C_0$ \cite{TV}, and the derivative $B^\prime_0=\partial B_0(a,b,c)/\partial a$. Explicit expressions for these functions are given in Appendix \ref{ap1}. Using the standard decomposition for the vector and tensor loop integrals \cite{TV} we obtain \begin{eqnarray} a_1 &=& -2+4C_{00}+2(C_{11}+C_1+r^2C_{22}+r^2C_2)+2(1- \hat{q}^2 +r^2)(C_{12}+C_0+C_1+C_2)\nonumber,\\ a_2 &=& 2r(C_1+C_2), \qquad a_3 = -4(C_{11}+C_{12}+C_1)-4r^2( C_{12}+C_{22}+C_2),\\ a_{wr}&=&\frac{1}{2}\Bigg[ 2-B_0(1,\hat{\mu}^2,1)-B_0(r^2,\hat{\mu}^2,r^2) +(1-\hat{\mu}^2)\Big(B_0(1,\hat{\mu}^2,1)-B_0(0,\hat{\mu}^2,1)\Big)\\ &&+\frac{(r^2-\hat{\mu}^2)}{ r^2} \Big(B_0(r^2,\hat{\mu}^2,r^2)-B_0(0,\hat{\mu}^2,r^2)\Big) +2(2+\hat{\mu}^2)B^\prime_0(1,\hat{\mu}^2,1)\nonumber \\ &&+2(2r^2+\hat{\mu}^2)B^\prime_0(r^2,\hat{\mu}^2,r^2)\Bigg] \nonumber . \end{eqnarray} Defining $f_1=1+r^2- \hat{q}^2 $ and $f_2=(f_1^2-4r^2)$ the coefficient functions take the form \begin{eqnarray} C_{00}&=&\frac{1}{4f_2}\Bigg[ f_2 +\hat{\mu}^2(f_1-2)B_0(1,1,\hat{\mu}^2)+\hat{\mu}^2(f_1-2r^2)B_0(r^2,r^2,\hat{\mu}^2)\nonumber\\ &&+(f_2+2 \hat{q}^2 \hat{\mu}^2) B_0( \hat{q}^2 ,1,r^2)+2\hat{\mu}^2(f_2+ \hat{q}^2 \hat{\mu}^2)C_0(1, \hat{q}^2 ,r^2,\hat{\mu}^2,1,r^2)\Bigg],\\ C_{11} &=& \frac{r^2}{f_2}+\frac{(f_1-2r^2)(1-r^2)}{2 \hat{q}^2 f_2}B_0(0,1,r^2) +\frac{f_1(\hat{\mu}^2-1)}{2f_2}B_0(0,1,\hat{\mu}^2)\\ &&+ \frac{3r^2\hat{\mu}^2(f_1-2r^2)}{f_2^2}B_0(r^2,r^2,\hat{\mu}^2)+ \frac{2 \hat{q}^2 \hat{\mu}^2(f_2+6 \hat{q}^2 r^2)-f_2(f_2+2 \hat{q}^2 r^2)}{2 \hat{q}^2 f_2^2}B_0( \hat{q}^2 ,1,r^2)\nonumber\\ &&+ \frac{\hat{\mu}^2 \Big( 6r^2(f_1-2)-f_2(f_1+2)\Big) } {2f_2^2}B_0(1,1,\hat{\mu}^2)\nonumber\\ &&+ \frac{2 \hat{\mu}^2 r^2 f_2+\hat{\mu}^4(f_2+6 \hat{q}^2 r^2) }{f_2^2}C_0(1, \hat{q}^2 ,r^2,\hat{\mu}^2,1,r^2)\nonumber,\\ C_{22}&=&\frac{1}{f_2}+\frac{(1-r^2)(2-f_1)}{2 \hat{q}^2 f_2}B_0(0,1,r^2) +\frac{2 \hat{q}^2 \hat{\mu}^2(f_2+6 \hat{q}^2 )-f_2(f_2+2 \hat{q}^2 )}{2 \hat{q}^2 f_2^2}B_0( \hat{q}^2 ,1,r^2)\nonumber\\ &&+\frac{(\hat{\mu}^2-r^2)f_1}{2r^2f_2}B_0(0,r^2,\hat{\mu}^2) +\frac{3\hat{\mu}^2(f_1-2)}{f_2^2}B_0(1,1,\hat{\mu}^2)\\ &&+\frac{\hat{\mu}^2\Big(6r^2(f_1-2r^2)-f_2(f_1+2r^2)\Big)}{2r^2f_2^2}B_0(r^2,r^2,\hat{\mu}^2)\nonumber\\ &&+\frac{\hat{\mu}^2\Big( 2f_2+\hat{\mu}^2(f_2+6 \hat{q}^2 )\Big)}{f_2^2}C_0(1, \hat{q}^2 ,r^2,\hat{\mu}^2,1,r^2) \nonumber ,\\ C_{12}&=&\frac{-f_1}{2f_2}+\frac{(1-r^2)(2r^2-f_1)}{2 \hat{q}^2 f_2}B_0(0,1,r^2) +\frac{r^2-\hat{\mu}^2}{f_2}B_0(0,r^2,\hat{\mu}^2)\nonumber\\ &&+\frac{\hat{\mu}^2\Big(6(f_1-2r^2)-f_2\Big)}{2f_2^2}B_0(1,1,\hat{\mu}^2) +\frac{\hat{\mu}^2\Big(6r^2(f_1-2)-f_2\Big)}{2f_2^2}B_0(r^2,r^2,\hat{\mu}^2)\\ &&+\frac{f_2(f_2+ \hat{q}^2 f_1)-2 \hat{\mu}^2 \hat{q}^2 (3 \hat{q}^2 f_1+f_2)}{2 \hat{q}^2 f_2^2}B_0( \hat{q}^2 ,1,r^2)\nonumber\\ &&+\frac{\hat{\mu}^2\Big(-f_1 f_2-\hat{\mu}^2(3 \hat{q}^2 f_1 +f_2) \Big)}{f_2^2}C_0(1, \hat{q}^2 ,r^2,\hat{\mu}^2,1,r^2)\nonumber,\\ C_1&=&\frac{1}{f_2}\Bigg[f_1 B_0(1,1,\hat{\mu}^2)+(2r^2-f_1)B_0( \hat{q}^2 ,1,r^2) -2r^2B_0(r^2,r^2,\hat{\mu}^2)\nonumber \\ &&+\hat{\mu}^2(2r^2-f_1)C_0(1, \hat{q}^2 ,r^2,\hat{\mu}^2,1,r^2)\Bigg],\\ C_{2}&=&\frac{1}{f_2}\Bigg[ -2 B_0(1,1,\hat{\mu}^2)+(2-f_1)B_0( \hat{q}^2 ,1,r^2)+f_1B_0(r^2,r^2,\hat{\mu}^2)\nonumber \\ &&+\hat{\mu}^2(2-f_1)C_0(1, \hat{q}^2 ,r^2,\hat{\mu}^2,1,r^2)\Bigg]. \end{eqnarray} The infinite parts of the regularized two-point functions can be shown to cancel in eq. (\ref{v1loop}). In the limit $\hat{\mu}\to 0$ the vertex correction diverges logarithmically. This divergence will be canceled by corresponding divergences in the bremsstrahlung contributions discussed in the next section. \section{Bremsstrahlung} The bremsstrahlung correction is found in the usual manner, by inserting a real gluon on the $c$ and $b$ quark lines. The calculation here is complicated by the four-body phase space with two massive final states. We follow the standard procedure of decomposing the four-body phase space into a two- and a three-body phase space by introducing the four-momentum $P=p_c+p_g$. In the rest frame of the $b$ quark this decomposition reads \begin{equation} dR_4 = dP^2\ dR_3(m_b; p_e, p_{\bar{\nu}}, P) dR_2(P; p_c, p_g). \end{equation} The ${\mathcal{O}}(\alpha_s)$ bremsstrahlung correction to the differential rate is given in terms of dimensionless variables ($\hat{P^0}=P^0/m_b$, $\hat{P}^2=P^2/m_b^2$) by \begin{eqnarray} {d \Gamma_{brems}^{(1)}(\hat{\mu}) \over {\rm d} y } &=& \alpha_s^{(V)}{G_F^2\, |V_{cb}|^2\, m_b^5 \over 192\ \pi^4 }\ \int d\hat{P}^2\ d\hat{P^0}\ (\hat{P^0}^2-\hat{P}^2)^{-5/2} \biggl[ 2 b_1 (1-2\hat{P^0}+\hat{P}^2) \nonumber \\ & & \qquad + b_2 (2-2\hat{P^0}-y) y + b_3 (1-y-\hat{P}^2) (2\hat{P^0}+y-\hat{P}^2-1) \nonumber \\ & & \qquad + b_4 (1-y-\hat{P}^2) y + b_5 (2\hat{P^0} +y-2) (1-2\hat{P^0}-y+\hat{P}^2) \biggl] . \label{r1loop} \end{eqnarray} For convenience the above rate has been written in terms of the coefficients $b_i$ \begin{eqnarray} b_1 &=& (\hat{P^0}^2-\hat{P}^2) \left( \hat{P}^2(c_2-c_1)+\hat{P^0}^2 c_1+c_3-\hat{P^0}(c_4+c_5) \right),\\ b_2 &=& (\hat{P^0}^2-\hat{P}^2)\hat{P}^2 c_1 +3\hat{P}^4 c_2+(\hat{P}^2+2\hat{P^0}^2) c_3 -3\hat{P^0}\hat{P}^2 (c_4+c_5) , \\ b_3 &=& (\hat{P^0}^2-\hat{P}^2) c_1 + (\hat{P}^2+2\hat{P^0}) c_2+3 c_3 - 3 \hat{P^0} (c_4+c_5) , \\ b_4 &=& -\hat{P^0}(\hat{P^0}^2-\hat{P}^2)c_1-3\hat{P^0}\hat{P}^2 c_2-3\hat{P^0} c_3+(\hat{P^0}^2+2\hat{P}^2) c_4+3\hat{P^0}^2 c_5,\\ b_5 &=& -\hat{P^0}(\hat{P^0}^2-\hat{P}^2) c_1-3\hat{P^0}\hat{P}^2 c_2-3\hat{P^0} c_3 + 3\hat{P^0}^2c_4 + (\hat{P^0}^2+2\hat{P}^2) c_5 , \end{eqnarray} which are linear combinations of \begin{eqnarray} c_1 &=& {4 (v_+^2 - v_-^2) \over h} + { 2 [ (h+\hat{\mu}^2-2\hat{P^0})^2+(\hat{\mu}^2-2\hat{P^0})^2+2 \hat{\mu}^2 (1+r^2) ] \over h} \ln( {2 v_+ -\hat{\mu}^2 \over 2 v_- - \hat{\mu}^2}) \nonumber \\ \quad &+& {4 (2+\hat{\mu}^2)(h-2\hat{P^0})(v_+-v_-) \over (2 v_+ -\hat{\mu}^2) (2 v_- -\hat{\mu}^2) } - {8 [(z+\hat{\mu}^2)(\hat{P^0}-h)+z\hat{P^0}] (v_+-v_-) \over h^2 } , \\ c_2 &=& {2 (v_+^2 - v_-^2) (2-h) \over h} - { [h \hat{\mu}^2(3\hat{\mu}^2-4\hat{P^0})+4\hat{\mu}^2(2\hat{P^0}-1)-16\hat{P^0}^2] \over h} \ln( {2 v_+ -\hat{\mu}^2 \over 2 v_- - \hat{\mu}^2}) \nonumber \\ \quad &+& {2 (\hat{\mu}^4-4)(2\hat{P^0}-\hat{\mu}^2)(v_+-v_-) \over (2 v_+ -\hat{\mu}^2) (2 v_- -\hat{\mu}^2) } - { 4 [h^2 (\hat{\mu}^2-\hat{P^0})+2\hat{P^0} (\hat{\mu}^2+2\hat{P}^2)] (v_+-v_-) \over h^2 } , \\ c_3 &=& {[(\hat{P}^2+r^2) (h^2+2(\hat{\mu}^2-2\hat{P^0})(h-2\hat{P^0})) -h\hat{\mu}^4 +4 r^2 \hat{P}^2 \hat{\mu}^2] \over h} \ln( {2 v_+ -\hat{\mu}^2 \over 2 v_- - \hat{\mu}^2}) \nonumber \\ \quad &+& {2 (2+\hat{\mu}^2) (\hat{\mu}^2-r^2-\hat{P}^2)(2\hat{P^0}-\hat{\mu}^2-h) (v_+-v_-) \over (2 v_+ -\hat{\mu}^2) (2 v_- -\hat{\mu}^2) } \nonumber \\ \quad &-& {4 [(\hat{P}^2+r^2)(h\hat{P^0}-h z+\hat{\mu}^2\hat{P^0})+4r^2\hat{P}^2\hat{P^0}] (v_+-v_-) \over h^2 } , \\ c_4 &=& {4 (\hat{P^0}-\hat{P}^2) (v_+^2 - v_-^2) \over h} + {2 (2+\hat{\mu}^2)(2\hat{P^0}-\hat{\mu}^2)(\hat{\mu}^2-2\hat{P^0}+h) (v_+-v_-)\over(2 v_+ -\hat{\mu}^2)(2 v_- -\hat{\mu}^2) } \nonumber \\ \quad &-& {2 [4\hat{P^0}^2(2\hat{P}^2+\hat{\mu}^2)+(\hat{\mu}^2-r^2) h (2\hat{P}^2+h-4\hat{P^0})+h\hat{P}^2(\hat{P}^2+r^2-8\hat{P^0})] (v_+-v_-) \over h^2 } \nonumber \\ \quad &-& {2 \over h} [2(\hat{\mu}^2-\hat{P^0})(h \hat{\mu}^2-2\hat{P^0} h+4\hat{P^0}^2)+h^2(1-\hat{P^0}+\hat{\mu}^2)- 2\hat{\mu}^2(r^2+r^2\hat{P^0}+\hat{P^0}) \nonumber \\ &+& \hat{\mu}^4(1+r^2-2\hat{P^0})] \ln( {2 v_+ -\hat{\mu}^2 \over 2 v_- - \hat{\mu}^2}) , \\ c_5 &=& {2 (\hat{\mu}^4-4) (\hat{P}^2+r^2-\hat{\mu}^2)(v_+-v_-)\over(2 v_+ -\hat{\mu}^2)(2 v_- -\hat{\mu}^2)} \nonumber \\ &-& {2 [\hat{\mu}^2h^2+(\hat{P}^2+r^2)(2\hat{\mu}^2+2h-h^2)+8r^2\hat{P}^2](v_+-v_-)\over h^2} \nonumber \\ &-& {2 [\hat{\mu}^4h-(\hat{P}^2+r^2)(\hat{\mu}^2h+4\hat{P^0})+2\hat{P}^2 \hat{\mu}^2] \over h} \ln( {2 v_+ -\hat{\mu}^2 \over 2 v_- - \hat{\mu}^2}) . \end{eqnarray} In the expressions for the $c_i$ we have put $h=\hat{P}^2-r^2$ and \begin{equation} v_\pm = { (\hat{P}^2+\hat{\mu}^2-r^2)\hat{P^0} \pm \sqrt{\hat{P^0}^2-\hat{P}^2} \sqrt{(\hat{P}^2+\hat{\mu}^2-r^2)^2-4 \hat{\mu}^2\hat{P}^2} \over 2\hat{P}^2 }. \end{equation} The integrals in eq. (\ref{r1loop}) are done numerically between the kinematic limits \begin{eqnarray} {(1-y)^2+\hat{P}^2 \over 2 (1-y)} \le &\hat{P^0}& \le {1+\hat{P}^2 \over 2} ,\\ (\hat{\mu} + r)^2 \le &\hat{P}^2& \le 1-y. \end{eqnarray} To improve the numerical stability for small $\hat{\mu}^2$ we found it useful to do the $\hat{P^0}$ integral with the variable $\ln(\hat{P^0}-r^2)$. The remaining limits for this four-body decay are \begin{equation} 0 \le \hat{\mu} \le \sqrt{1-y} - r, \qquad 0 \le y \le 1-r^2. \end{equation} \section{The $\alpha_s^2\beta_0$ correction} Combining the corrections from virtual and real gluon radiation, eqs. (\ref{v1loop},\ref{r1loop}), we obtain \begin{equation}\label{glue} \frac{ {\rm d} \Gamma^{(1)}(\hat{\mu})}{ {\rm d} y} = \frac{ {\rm d} \Gamma^{(1)}_{virt}(\hat{\mu})}{ {\rm d} y} +\frac{ {\rm d} \Gamma^{(1)}_{brems}(\hat{\mu})}{ {\rm d} y}\Theta(\sqrt{1-y}-r-\hat{\mu}). \end{equation} In the $\hat{\mu}\to 0$ limit eq. (\ref{glue}) yields the one-loop correction to the electron spectrum. We have checked that our expression reproduces the result in \cite{jk} in this limit. The $\alpha_s^2\beta_0$ part of the two-loop correction is related to the one-loop expression calculated with a massive gluon by \cite{sv} \begin{equation} \frac{ {\rm d} \Gamma^{(2)}}{ {\rm d} y}=-\frac{{\alpha_s^{(V)}}\beta_0}{4\pi} \int\limits_{0}^{\infty}\frac{ {\rm d} \hat{\mu}^2}{\hat{\mu}^2} \left( \frac{ {\rm d} \Gamma^{(1)}(\hat{\mu})}{ {\rm d} y}- \frac{1}{1+\hat{\mu}^2}\frac{ {\rm d} \Gamma^{(1)}(0)}{ {\rm d} y}\right). \end{equation} Note that $\alpha_s^{(V)}$, defined in the V-scheme of BLM \cite{BLM}, is related to the more familiar $\bar\alpha_s$ defined in the $\overline{\rm MS}$ scheme by \begin{equation} \alpha_s^{(V)}=\bar\alpha_s+\frac{5}{3}\frac{\bar\alpha_s^2}{4\pi}\beta_0+\cdots. \end{equation} $\alpha_s$ is evaluated at $m_b$ unless stated otherwise. In the $\overline{\rm MS}$ scheme the $\bar\alpha_s^2\beta_0$ part of the two-loop correction reads \begin{equation} \frac{ {\rm d} \Gamma^{(2)}}{ {\rm d} y}=\frac{5}{3}\frac{\bar\alpha_s\beta_0}{4\pi} \frac{ {\rm d} \Gamma^{(1)}(0)}{ {\rm d} y} -\frac{\bar\alpha_s\beta_0}{4\pi}\int\limits_{0}^{\infty}\frac{ {\rm d} \hat{\mu}^2}{\hat{\mu}^2} \left( \frac{ {\rm d} \Gamma^{(1)}(\hat{\mu})}{ {\rm d} y}- \frac{1}{1+\hat{\mu}^2}\frac{ {\rm d} \Gamma^{(1)}(0)}{ {\rm d} y}\right). \end{equation} The dispersion integral has to be done with some care. We found that using $\ln(\hat{\mu}^2)$ instead of $\hat{\mu}^2$ as the integration variable simplifies the numerical evaluation considerably. In Fig. 1 we plot the $\bar\alpha_s^2\beta_0$ part of the two-loop correction\footnote{ In Appendix \ref{ap2} we give a polynomial fit to our results that reproduces the numerical results to better than 1\% for mass ratios in the range $0.29\le r\le 0.37$.} and for comparison the one-loop correction to the electron spectrum, using $r=0.29$, $\bar\alpha_s=0.2$, $n_f=3$, and dividing by $\Gamma_0=G_F^2|V_{cb}|^2m_b^5/192\pi^3$. Except for electron energies close to the endpoint, the $\alpha_s^2\beta_0$ corrections are about half as big as the first order corrections. The perturbation series appears to be controlled but the higher order corrections clearly are not negligible. Integrating over the electron energy we reproduce the result for the correction to the total rate given in Ref. \cite{asmark}. In \cite{gklw} the HQET matrix elements $\bar\Lambda,\lambda_1$ were extracted from the lepton spectrum using the experimentally accessible observables \begin{equation} R_1=\frac{\int_{1.5{\rm GeV}} {\rm d} E_\ell \, E_\ell {\rm d} \Gamma/ {\rm d} E_\ell} {\int_{1.5{\rm GeV}} {\rm d} E_\ell \, {\rm d} \Gamma/ {\rm d} E_\ell}, \qquad R_2=\frac{\int_{1.7{\rm GeV}} {\rm d} E_\ell \, {\rm d} \Gamma/ {\rm d} E_\ell} {\int_{1.5{\rm GeV}} {\rm d} E_\ell \, {\rm d} \Gamma/ {\rm d} E_\ell}. \end{equation} It is straightforward to calculate the $\bar\alpha_s^2\beta_0$ corrections to these quantities. In the spirit of HQET, we use the spin averaged meson masses $\overline{m}_B=5.314{\rm GeV}$, and $\overline{m}_D=1.975{\rm GeV}$ instead of quark masses. Keeping only two-loop corrections that are proportional to $\beta_0$, and neglecting terms of order $\bar\alpha_s^2\beta_0\Lambda_{QCD}/\overline{m}_B$ we find \begin{eqnarray}\label{r1r2} R_1&=&1.8059-0.035\frac{\bar\alpha_s}{\pi}-0.082\frac{\bar\alpha_s^2\beta_0}{\pi^2}+\cdots,\\ R_2&=&0.6581-0.039\frac{\bar\alpha_s}{\pi}-0.098\frac{\bar\alpha_s^2\beta_0}{\pi^2}+\cdots\nonumber, \end{eqnarray} where the ellipsis denote the other contributions including nonperturbative corrections discussed in \cite{gk,gklw}. The BLM scales for these quantities are $\mu_{BLM}(R_1) = 0.01 \overline{m}_B$, and $\mu_{BLM}(R_2) = 0.007\overline{m}_B,$ reflecting the fact that the second order corrections are larger than the first order. This is a result of the almost complete cancellation of the first order perturbative corrections from the denominators and numerators in $R_{1,2}$. In eq.(\ref{r1r2}) the BLM scales for the numerators and denominators are separately comparable to the BLM scale for the total rate $\mu_{BLM}\approx 0.1\overline{m}_B$. Therefore the very low BLM scales of $R_{1,2}$ do not necessarily indicate badly behaved perturbative series. In order to demonstrate the impact of the $\bar\alpha_s^2\beta_0$ corrections on the extraction of $\bar\Lambda,\lambda_1$, we repeat the analysis of \cite{gklw} neglecting nonperturbative corrections of order $(\Lambda_{QCD}/m_b)^3$ which may be substantial \cite{gk}. Because of the higher order nonperturbative corrections, large theoretical uncertainties have to be assigned to the extracted values of $\bar\Lambda,\lambda_1$. We find that the central values are moved from $\bar\Lambda=0.39\pm0.11{\rm GeV},\,\lambda_1=-0.19\pm 0.10 {\rm GeV}^2$ to $\bar\Lambda=0.33{\rm GeV},\,\lambda_1=-0.17$. The shift in the values of the HQET matrix elements lies well within the $1\sigma$ statistical error of the previously extracted values. However, using the values of the HQET matrix elements extracted at a given order in $\alpha_s$ to predict physical observables at the same order in $\alpha_s$, guarantees that the renormalon ambiguity in $\bar\Lambda$ and $\lambda_1$ will cancel\cite{renorm,renorm2} if the expansion is continued to sufficiently high orders in $\alpha_s$. Thus including the $\bar\alpha_s^2\beta_0$ parts in the determination of $\bar\Lambda,\lambda_1$ allows one to calculate the $\overline{\rm MS}$ quark masses consistently at order $\bar\alpha_s^2\beta_0$. To second order in $\Lambda_{QCD}/m_q$ and to order $\bar\alpha_s^2\beta_0$ we have \begin{equation} \overline{m}_q(m_q)=\left(\overline{m}_{Meson}- \bar\Lambda+\frac{\lambda_1}{2m_q}+\cdots\right) \left(1-\frac{4\bar\alpha_s(m_q)}{3\pi} -1.56\frac{\bar\alpha_s^2(m_q)\beta_0}{\pi^2}+\cdots\right), \end{equation} where $m_q$ is the $b$ or $c$ quark pole mass and $\overline{m}_{Meson}$ is the corresponding spin averaged meson mass. With $\bar\alpha_s(m_b)=0.22,\,\bar\alpha_s(m_c)=0.39$ this yields $\overline{m}_b(m_b)=4.16{\rm GeV}, \overline{m}_c(m_c)=0.99{\rm GeV}$ for the $\overline{\rm MS}$ quark masses, albeit with large theoretical uncertainties due to the effect of the higher order nonperturbative corrections on the extraction of $\bar\Lambda,\lambda_1$\cite{gk}. The value of $\overline{m}_b(m_b)$ is in good agreement with lattice calculations $\overline{m}_b(m_b)=4.17\pm 0.06{\rm GeV}$ and $\overline{m}_b(m_b)= 4.0\pm 0.01{\rm GeV}$\cite{lat}. The weak mixing angle $|V_{cb}|$ can be determined by comparing the theoretical prediction for the total rate with experimental measurements. Including all corrections discussed in \cite{gklw} we find at order $\alpha_s^2\beta_0$ \begin{equation} |V_{cb}|=0.043\left(\frac{ Br(B\to X_c\ell\bar{\nu_{\ell}})}{0.105} \frac{ 1.55{\rm ps}}{\tau_B}\right)^{1/2}. \end{equation} \section{Conclusions} We have calculated the ${\cal O}(\alpha_s^2\beta_0)$ corrections to the electron spectrum in $b\to c\ell\bar{\nu_{\ell}}$ decays which turn out to be rather large, about 50\% of the one-loop corrections. These corrections can be included in the extraction of the HQET matrix elements $\bar\Lambda,\lambda_1$. We obtain $\bar\Lambda=0.33{\rm GeV}$ and $\lambda_1=-0.17{\rm GeV}^2$, both somewhat lower than the values extracted at ${\cal O}(\alpha_s)$. Using these values and including ${\cal O}(\alpha_s^2\beta_0)$ corrections we obtain $\overline{m}_b(m_b)=4.16{\rm GeV},\, \overline{m}_c(m_c)=0.99{\rm GeV}$ for the $\overline{\rm MS}$ quark masses and $|V_{cb}|=0.043( Br(B\to X_c\ell\bar{\nu_{\ell}})/0.105\times 1.55{\rm ps}/\tau_B)^{1/2}$. These results have large theoretical uncertainties due to the effect of nonperturbative corrections of order $(\Lambda_{QCD}/m_b)^3$ on the extraction of $\bar\Lambda,\lambda_1$. \acknowledgments We would like to thank Anton Kapustin, Zoltan Ligeti and Mark Wise for helpful discussions. This work was supported in part by the Department of Energy under grant DE-FG03-92-ER 40701.
proofpile-arXiv_065-401
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} The evolution of living organisms is a fascinating phenomenon that has intrigued the imagination of the scientific and non-scientific community. However, the formulation of mathematical models falls necessarily to drastic simplifications. For example, evolution has often been considered as a ``walk'' in a rugged landscape. Following this line, Bak and Sneppen (BS) have proposed a model of biological evolution \cite{bs} that has become quite interesting to the physics community due to its simplicity and the new insight it provides to the problem. It has been shown that this model evolves to a self-organized critical state (SOC), and is kept there by the means of avalanches of evolutionary activity. This is appealing for a model of biological evolution, since it has been observed that life on Earth could be in a SOC state \cite{kauf,newm}. Nevertheless, models based in fitness landscapes, or in a concept of fitness different from the biological one, have been criticized from a biological point of view \cite{newm,jong}. Since one of the characterizing aspects of life, and perhaps the most fundamental one, is that of self-replication, it is our belief that more realistic models should involve a dynamic population for each species. The starting point of combining population dynamics with evolution is the association of the rates of birth and death and the carrying capacity with phenotypes (observable features that arise from the genotype and are, then, subject to mutation) \cite{roug}. The fitness, namely the expected number of offsprings produced by an individual, arise from them. In this way, the process of natural selection is directed by the ecological interactions instead of by a non-biological notion of relative fitness. Extinction is an essential component of evolution. The great majority of species that have ever lived on Earth are now extinct\cite{raup}. There exist competing hypothesis that account extinction as originating from within the biosystem, or from external causes --what has been called ``bad genes or bad luck''. In any case, the pattern of extinctions and of surviving species or groups of species is certainly an interesting problem to model, to understand, and eventually to check with the fossil record. We show in this contribution a simple model of a large ecological system in evolution. This produces features of extinction similar to those claimed for the biosystem on Earth. We have chosen to study an ecological model in which each species consists of a population interacting with the others, that reproduces and evolves in time. The system is supposed to be a food chain, and the interactions to be predator-prey. Mutations that change the interactions are supposed to occur randomly at a low rate. Extinctions of populations result from the predator-prey dynamics. This approach can be thought as middle way between the microscopic simulation of ``artificial life'' by Ray and others \cite{ray}, and the coarse-grain description of models like BS's. \section{The model} Our model ecosystem consists of a number of species that interact and evolve in time. In the course of its time evolution the populations grow and shrink following a set of equations. Eventually, some of the species become extinct as a result of their interaction with the others. Every now and then we change one of the phenotypic features of one of the species, mimicking a random mutation of its genome. This modification produces a perturbation in the dynamics of the ecosystem, and eventually leads to the extinctions. To be more precise, let's consider a simple example of a food web, namely a one-dimensional food chain. $N$ species interact in such a way that the species $i$ feeds on the species $i-1$, and is eaten by the species $i+1$. The species $1$ is an autotroph: it feeds at a constant rate on an ``environment''. The species $N$, the top of the chain, is not eaten by any species, but dies giving its mass to the environment. Each species has a population that evolves in time and interacts with its neighbors in the chain. Furthermore, we consider this evolution in discrete time, which is often more realistic than a continuous one \cite{roug} and simpler to simulate in a computer. As has been said above, each species acts as a predator with respect to the one preceding it in the chain, and as a prey with respect to the one following it. As a further simplification, we suppose that there are no intrinsic birth and death rates, apart from those arising from the predation and prey contributions. Let's propose the equations governing this behaviors \cite{murray}. As a predator, the ``population'' (a continuous density) of the species $i$, $n^i$, changes from time $t$ to time $t+1$ according to: \begin{equation} \Delta n_t^i=k_in_t^{i-1}n_t^i\left( 1-n_t^i/c_i\right) , \label{predator} \end{equation} where $k_i$ is a rate of growth of the predator population and $c_i$ is a carrying capacity that accounts for a limitation imposed by the environment. Note that (\ref{predator}) includes this carrying capacity in a logistic factor to avoid an unbounded growth of the population. Also observe that the growth is proportional to the population of preys, without a ``satiation'' factor. Similarly, as a prey, the population of the species $i$ will diminish according to: \begin{equation} \Delta n_t^i=-g_in_t^in_t^{i+1}. \end{equation} The parameters $k_i$, $g_i$ and $c_i$ are the phenotypic features of our species. In the course of the evolution we allow them to change, mimicking random mutations. Moreover, they are the same for all the individuals of each single population. We do not model races, traits, polymorphisms or any phenotypic variation within a species, and when a mutation occurs it is assumed that the whole population ``moves'' instantly to the new state. In this sense, we are modelling the co-evolution of the species and disregarding the evolution of a single one as well as other important phenomena like the formation of new species \cite{vand}. Combining the two roles of predator and prey that each species performs, and the special status of the ends of the chain, we can write the following set of equations for the evolution of the system: \begin{equation} \left\{ \begin{array}{lcl} \Delta n_t^i & = & k_in_t^{i-1}n_t^i\left( 1-n_t^i/c_i\right) -g_in_t^{i+1}n_t^i \mbox{\ \ \ \ for $i=1 \dots N$} \\ n^0 & = & n^{N+1} = 1, \end{array} \right. \label{model} \end{equation} where we have introduced two fictitious species, $0$ and $N+1$, to take account for the border condition. We make two simplifications to the system (\ref{model}): 1) we assume that all the carrying capacities are equal, and equal to $1$; 2) we assume that $g_i=k_{i+1}$. In this way we reduce the number of parameters that define the phenotypic features of the ecosystem. The dynamics of the system is as follows. At time $t=0$, all the populations and interactions are chosen at random with uniform distribution in the interval $\left( 0,1\right) $. Then the populations begin to evolve according to the system (\ref{model}). In the course of the evolution driven by eq.(\ref{model}) a population can go to zero. As this can happen asymptotically, we consider a species extinct if its population drops below a given threshold. This is reasonable since actual biological populations are discrete. In order to keep constant the number of species we replace an extinct one with a new one, which can be thought as a species coming to occupy the niche left by the extinct one \cite{footnote}. The new population, and the (two) new interactions with its neighbors in the chain, are also drawn at random from a uniform distribution in the interval $\left( 0,1\right) $. On top of this dynamics of predation and extinctions, we introduce random mutations. In each time step a mutation is produced with probability $p$; the species to mutate is chosen at random and the mutation itself consists of the replacement of the species with a new one, with a new population and new interactions with its predators and preys (all random in $(0,1)$). Observe that we do not introduce the fitness of a species as a dynamical variable. We do not even need to compute it from the ``phenotypes'' $k_i$. The fitness, the degree of adaptation of a species to the ecosystem, arises from the phenotypes, the populations, and the dynamics, and it determines whether a species will thrive or become extinct. Chance is introduced by the random mutations (and the random replacement of extinct species). It provides the material the natural selection works on. This, in turn, determines the survival of the fittest by simply eliminating from the system those species that cannot cope with the competing environment. We believe in this way we avoid a fundamental problem in the models of evolution as a walk in a fitness landscape, namely that the concepts of fitness is not the biological one \cite{jong}. \section{Results of the numerical simulation} We have run our model for several chain sizes, ranging from $50$ to $1000$ sites, and for times of about $10^7$ steps. In the results reported below we let the system evolve, during a transient period, from the initial random state to an organized one. In fig.\ \ref{pob} we show a typical evolution of the whole population, $% \sum_1^N n^i$. Although each population greatly changes in the course of time (what is not shown in the picture), we observe that the whole population remains relatively stable. This is due to the saturation factor in the predation term of the evolution equations. This whole population shows a short time oscillatory dynamics governed by the competition between species through the set of equations (% \ref{model}), and a long time evolution characterized by periods of relative stasis and periods of fast change. This feature is the effect of mutations and extinction of some species. Without the extinctions and mutations, the dynamics of the system should probably be chaotic. But it is not this feature that we want to analyze here. Instead, we shall focus on the pattern of extinctions. As the set of $k_i$ represents the phenotypes of the whole ecosystem, its distribution, $P(k)$, can be used to characterize its state. Let's observe what happens in the course of time, including the transient mentioned above. Initially the $k_i$ are chosen at random, and thus its distribution is flat in $(0,1)$, with mean $0.5$. This is shown in fig.\ \ref{k} as a full line. As time passes, and as a result of the dynamics, this distribution shifts to a non-uniform one, as shown in fig.\ \ref{k} with dashed lines. The whole distribution shifts towards lower values of the interaction, showing a tendency of the system to reduce the coupling between the species. In the course of the evolution this distribution fluctuates following the pattern of mutations and extinctions, but preserves its form. Fig.\ \ref{kmed} shows the above mentioned fluctuations in $P(k)$ as the evolution of the mean value of $k_i$ in the system, after the transient. It corresponds to the same run as fig.\ \ref{pob}, and the same time window is shown. Similarly to that, it displays a pattern of periods of stasis interrupted by periods of fast change, but without the short time oscillations displayed by the population. There are periods of stasis of all lengths, to a degree that the unique scale of the figure cannot display. This feature of a lack of a typical length will be analyzed immediately. Observe in this figure that the mean value oscillates around $0.24$, corresponding to a distribution like that shown in fig.\ \ref{k} with a dotted line. The extinction events also display this characteristic pattern of periods of stasis and periods of change, without a typical size. In order to characterize this, we have chosen the time between two consecutive extinction events, $\tau$, which distribution is shown in fig.\ \ref{ext} for several system sizes an probabilities of mutation. Observe that they follow a power law for several decades of large values of $\tau$, before a region where the effects of the finite size of the system start to appear. This is a sign that the system has self-organized into a critical state. In other words, the extinction events are distributed in the time axis in such a way that the time between extinctions does not have a characteristic duration --as should have if the distribution were exponential. Extinctions appear to come in bursts, or avalanches, of any size. In fig.\ \ref{aval} the pattern of extinction events of the system is seen in the course of time. The graphic displays time in the abscissa and the index in the food chain in the ordinate. Each dot marks the moment in which a species has become extinct. Each cross, a species that suffers a chance mutation. It can be seen that some mutations trigger avalanches of extinction, and that these propagate in the ``prey'' direction. (Bear in mind that an extinct species is replaced by a random new one, most probably with a larger population than its predecessor, and observe that this has a negative impact in the corresponding {\em prey}.) It is also apparent that this avalanches have a complex shape in space-time. It is not easy to measure their size since, as can be seen in fig.\ \ref{aval}, they overlap. See, for example, a mutation that is {\em not} followed by any avalanche (lower left), another that triggers a very small one (lower right), and several that start events of varying size. In any case, let's define a time step, $\Delta t$, divide the time axis with it, and count the number of extinctions in each interval. Now, let's call the fraction of species that have become extinct in each interval the {\em size}, $S$, of the extinction. $S$ will obviously depend on the time step and on the size of the system: $S=S(\Delta t,N)$. If the system is in a critical state this function will obey some scaling law on the variable $N$. In fig.\ \ref{scaling} we have scaled the distribution of the system size $S(N)$, $P(S,N)$, obtained for different system sizes according to the ansatz: \begin{equation} P(S,N)=N^\beta f(S\cdot N^\nu ). \end{equation} We can observe that the four curves collapse to a single one, showing the scaling behaviour that is typical of a critical state. \section{Conclusions} We have introduced a simple model of co-evolution and extinction in a food chain. This consists of a finite chain of species of predators and preys. Their populations evolve in time following Lotka-Volterra-like equations. Evolution is mimicked by randomly changing a phenotype. Natural selection is provided by the deterministic behaviour of the dynamical system, that produces the extinction of any species that cannot cope with its interactions. No relative fitness or fitness landscape had to be invoked. Nevertheless, the pattern of extinctions displayed by this toy ecosystem appears to be similar to that proposed for the biosystem on Earth. Namely, the system seems to be in a critical state, in which extinctions occur in avalanches. The time between extinctions, and the lifetime of any species follow distributions that behave like power-laws of time, implying that there is no typical size for the time that a species remains in the system. I should be of interest, in a future work, to study the precise instability that produces the shift of the distribution of the interactions towards low values. The analytical treatment of this problem is currently under study. The author greatly acknowledges Ruben Weht for invaluable discussions, and thank Hilda Cerdeira for a careful reading of the manuscript.
proofpile-arXiv_065-402
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} Since the previous HERA workshop in 1991 significant progress has been made on the theoretical side in understanding the production of heavy quarks in electron proton collisions. Improvements in available experimental techniques and particularly the expected increase in luminosity amply justify this effort. In general the progress consists of the calculation of all $O(\alpha_s)$ corrections to the processes of interest, thus improving the accuracy of the theoretical predictions both in shape and normalization. At the time of the previous workshop the only NLO calculations available were for the case of inclusive photoproduction \cite{photoincl}. In the meantime NLO calculations have also been performed for inclusive electroproduction \cite{LRSNF,LRSNH,LRSND}, and both have been extended to the fully differential cases \cite{FMNR,HS,H}. Therefore, meaningful and extensive comparisons between theory and data can now be made. In what follows we review how the deeply inelastic electroproduction process allows us to explore, in detail, three areas of perturbative QCD in particular. We first discuss the inclusive case, via the structure function $F_2(x,Q^2,m^2)$. We show that this structure function for the case of charm suffers from only very modest theoretical uncertainty, that its NLO corrections are not too large, and that it is sensitive to the shape of the small-$x$ gluon density. Next we treat single particle differential distributions in the charm kinematical variables, and also charm-anticharm correlations. Because many distributions can be studied, many QCD tests can be performed. Examples are tests of the production mechanism (boson-gluon fusion), studies of gluon radiation patterns, and dependence on scales such as deep-inelastic momentum transfer $Q$, the heavy quark mass $m$ (with enough luminosity one can detect a sizable sample of bottom quarks), the transverse momentum of the charm quark, etc. Finally, in the last section, we review the theoretical status of the boson-gluon fusion description of charm production at small and very large $Q$. In essence, it involves answering the question: when is charm a parton? \section{Structure Functions and Gluon Density} This section has some overlap with the more detailed review on heavy flavour structure functions in the structure function section. Here we only present the most salient features. The reaction under study is \begin{equation} e^-(p_e) + P(p) \rightarrow e^-(p_e') + Q(p_1)(\bar{Q}(p_1))+X\,,\label{one} \end{equation} where $P(p)$ is a proton with momentum $p$, $Q(p_1)(\bar{Q}(p_1))$ is a heavy (anti)-quark with momentum $p_1$ ($p_1^2 = m^2$) and $X$ is any hadronic state allowed. Its cross section may be expressed as \begin{equation} \label{three} \frac{d^2\sigma}{dxdQ^2} = \frac{2\pi\alpha^2}{x\,Q^4} \left[ ( 1 + (1-y)^2 ) F_2(x,Q^2,m^2) -y^2 F_L(x,Q^2,m^2) \right]\,, \end{equation} where \begin{equation} q = p_e - p_e'\,, \qquad Q^2 = -q^2\,,\qquad x = \frac{Q^2}{2p\cdot q} \,, \qquad y = \frac{p\cdot q}{p \cdot p_e}\,. \label{four} \end{equation} The inclusive structure functions $F_2$ and $F_L$ were calculated to next-to-leading order (NLO) in Ref.~\cite{LRSNF}. The results can be written as \begin{eqnarray} F_{k}(x,Q^2,m^2) &=& \frac{Q^2 \alpha_s}{4\pi^2 m^2} \int_x^{z_{\rm max}} \frac{dz}{z} \Big[ \,e_H^2 f_g(\frac{x}{z},\mu^2) c^{(0)}_{k,g} \,\Big] \nonumber \\&& +\frac{Q^2 \alpha_s^2}{\pi m^2} \int_x^{z_{\rm max}} \frac{dz}{z} \Big[ \,e_H^2 f_g(\frac{x}{z},\mu^2) (c^{(1)}_{k,g} + \bar c^{(1)}_{k,g} \ln \frac{\mu^2}{m^2}) \nonumber \\ && +\sum_{i=q,\bar q} \Big[ e_H^2\,f_i(\frac{x}{z},\mu^2) (c^{(1)}_{k,i} + \bar c^{(1)}_{k,i} \ln \frac{\mu^2}{m^2}) + e^2_{L,i}\, f_i(\frac{x}{z},\mu^2) d^{(1)}_{k,i} \, \Big] \,\Big] \,, \label{strfns} \end{eqnarray} where $k = 2,L$ and the upper boundary on the integration is given by $z_{\rm max} = Q^2/(Q^2+4m^2)$. The functions $f_i(x,\mu^2)\,, (i=g,q,\bar q)$ denote the parton densities in the proton and $\mu$ stands for the mass factorization scale, which has been put equal to the renormalization scale. The $c^{(l)}_{k,i}(\eta, \xi)\,,\bar c^{(l)}_{k,i} (\eta, \xi)\,, (i=g\,,q\,,\bar q\,;l=0,1)$ and $d^{(l)}_{k,i}(\eta, \xi)$, $(i=q\,,\bar q\,;l=0,1)$ are coefficient functions and are represented in the $\overline{\rm MS}$ scheme. They depend on the scaling variables $\eta$ and $\xi$ defined by \begin{equation} \eta = \frac{s}{4m^2} - 1\quad \qquad \xi = \frac{Q^2}{m^2}\,. \end{equation} where $s$ is the square of the c.m. energy of the virtual photon-parton subprocess which implies that in (\ref{strfns}) $z=Q^2/(Q^2+s)$. In eq.~(\ref{strfns}) we distinguished between the coefficient functions with respect to their origin. The coefficient functions indicated by $c^{(l)}_{k,i}(\eta, \xi),\bar c^{(l)}_{k,i}(\eta, \xi)$ originate from the partonic subprocesses where the virtual photon is coupled to the heavy quark, whereas the quantity $d^{(l)}_{k,i}(\eta, \xi)$ comes from the subprocess where the virtual photon interacts with the light quark. Hence the former are multiplied by the charge squared of the heavy quark $e_H^2$, and the latter by the charge squared of the light quark $e_L^2$ respectively (both in units of $e$). Terms proportional to $e_H e_L$ integrate to zero for the inclusive structure functions. Furthermore we have isolated the factorization scale dependent logarithm $\ln(\mu^2/m^2)$. A fast program using fits to the coefficient functions \cite{RSN} is available. The first thing to note about eq.~(\ref{strfns}) is that the lowest order term contains only the gluon density. Light quark densities only come in at next order, and this is the reason $F_2(x,Q^2,m^2)$ is promising as a gluon probe. To judge its use as such, we must examine some of the characteristics of this observable. These are: the size of the $O(\alpha_s)$ corrections, the scale dependence, the mass dependence, its sensitivity to different gluon densities, and the relative size of the light quark contribution. These are the issues we investigate in this section. We take the charm mass 1.5 GeV, the bottom mass 5 GeV, the factorization scale equal to $\sqrt{Q^2+m^2}$ and choose at NLO the CTEQ4M \cite{cteq4} set of parton densities, with a two-loop running coupling constant for five flavors and $\Lambda = 202 $ MeV, and at LO the corresponding CTEQ4L set, with a one-loop running coupling with five flavors and $\Lambda = 181 $ MeV. \begin{figure}[htbp] \hspace*{1.5cm} \epsfig{file=fig1.ps,bbllx=0pt,bblly=20pt,bburx=575pt,bbury=880pt,% width=10cm,angle=-90} \caption[junk]{{\it $F_2(x,Q^2,m^2)$ vs. $x$ at LO and NLO for two values of $Q^2$. The shaded areas indicate the uncertainty due to varying the charm mass from 1.3 to 1.7 GeV. }} \label{FIGF2C1} \end{figure} In Fig.~\ref{FIGF2C1} we display $F_2(x,Q^2,m^2)$ vs. $x$ for two values of $Q^2$ at LO and NLO. The scale dependence is much reduced by including the NLO corrections (when varying $\mu$ from $2$ to $1/2$ times the default choice, the structure function varies from, at LO, at most 20\% and 13\% at $Q^2 = 10$ and $50$ GeV$^2$ respectively, to at most 5\% and 3\% at NLO), but the dominant uncertainty is due to the charm mass and stays roughly constant, amounting at NLO maximally to about 16\% for $Q^2=10$ GeV$^2$ and 10\% for $Q^2=50$ GeV$^2$. The feature that the LO result is mostly larger than the NLO ones is due to the use of LO parton densities and one-loop $\alpha_s$, and scale choice. Had we used NLO densities and a two-loop $\alpha_s$, or chosen the scale $\mu$ equal to $m$, the LO result would have been below the NLO result. In the first case the size of the corrections is then about 40\% at the central values at $Q^2 = 10$ GeV$^2$, and 25\% at $Q^2 = 50$ GeV$^2$, and in the second case, at small $x$, about 20\% and 30\% respectively. \begin{figure}[htbp] \hspace*{1.5cm} \epsfig{file=fig2.ps,bbllx=0pt,bblly=20pt,bburx=575pt,bbury=880pt,% width=10cm,angle=-90} \caption[junk]{{\it $F_2(x,Q^2,m^2)$ vs. $x$ at NLO for two choices of parton densities. The shaded areas again indicate the uncertainty due to varying the charm mass from 1.3 to 1.7 GeV. }} \label{FIGF2C2} \end{figure} In the next figure, Fig.~\ref{FIGF2C2}, we show for the same values of $Q^2$ an important property, namely the sensitivity of the NLO $F_2$ to different parton density parametrizations. In this case we compare the CTEQ2MF set \cite{cteq2}, whose gluon density stays quite flat when $x$ becomes small, and the GRV94 set \cite{GRV94}, which has a steeply rising gluon density. One sees that the difference is visible in the structure function. Finally we remark that the contribution of light quarks to the charm structure function is typically less than 5\%. The bottom quark structure function is suppressed by electric charge and phase space effects and amounts to less than 2\% (5\%) at $Q^2=10\,(50)$ GeV$^2$ of the charm structure function. Previous investigations of the scale and parton density dependence of $F_2$ using the same NLO computer codes are available in \cite{vogt} and \cite{grs}. We conclude that $F_2(x,Q^2,m^2)$ for charm production is an excellent probe to infer the gluon density in the proton at small $x$. The NLO theoretical prediction suffers from fairly little uncertainty, and the QCD corrections are not too large. See the section on structure functions in these proceedings for many more details, where also a comparison with (preliminary) data is shown. Therefore in view of a large integrated luminosity, a theoretically well-behaved observable, and promising initial experimental studies \cite{H1paper,zeus} a precise measurement at HERA of the gluon density should be possible. \section{Single Particle Distributions and Heavy Quark Correlations} In this section we leave the fully inclusive case and examine in more detail the structure of the final state of the reaction \begin{equation} e^-(p_e) + P(p) \rightarrow e^-(p_e') + Q(p_1)+\bar{Q}(p_2)+X\,. \label{react2} \end{equation} By studying various differential distributions of the heavy quarks we can learn more about the dynamics of the production process than from the structure function alone. Single particle distributions $dF_2(x,Q^2,m^2,v)/dv$, where $v$ is the transverse momentum $p_T$ or rapidity $y$ of the charm quark, were presented in NLO in \cite{LRSND} for various choices of $x$ and $Q^2$. The LO distributions differed significantly from the NLO ones, so that the effect of $O(\alpha_s)$ corrections on such distributions cannot be described by a simple K-factor. The $O(\alpha_s)$ corrections to $F_k(x,Q^2,m^2)$ in a fully differential form were calculated in Ref.~\cite{HS} using the subtraction method. Recently \cite{H}, these fully differential structure functions were incorporated in a Monte-Carlo style program resulting in the $O(\alpha_s)$ corrections for reaction (\ref{react2}). The program for the full cross section, generated according to Eq.\ (\ref{three}), allows one to study correlations in the lab frame. The phase space integration is done numerically. Therefore, it is possible to implement experimental cuts. It furthermore allows the use of a Peterson type fragmentation function. For details about the calculational techniques we refer to Ref.~\cite{HS,H}. Here we show mainly results. \begin{figure}[htbp] \hspace*{.5cm} \epsfig{file=fig3.ps,bbllx=0pt,bblly=200pt,bburx=575pt,bbury=600pt,% width=13cm,angle=0} \caption[junk]{{\it Differential cross sections and ZEUS data. }} \label{FIGF2C3} \end{figure} Shown in Fig.~\ref{FIGF2C3} are various distributions $d\sigma/dv$ for the reaction (\ref{react2}), where the heavy (anti)quark has fragmented into a $D^*$ meson, with $v$ representing (a) the $D^*$ transverse momentum $p_T^{{D^*}}$ (b) its pseudorapidity $\eta^{{D^*}}$ (c) the hadronic final state invariant mass $W$ (d) $Q^2$ for the kinematic range 5 GeV$^2 < Q^2 < $ 100 GeV$^2$, $0<y<0.7$, $1.3\,{\rm GeV} < p_T^{D^*} < 9 {\rm GeV}$ and $|\eta^{{D^*}}| < 1.5$. The data are from a recent ZEUS analysis \cite{zeus}. The NLO theory curves have been produced by using the GRV \cite{GRV94} parton density set, with Peterson fragmentation \cite{peterson}. The dashed line is for $\mu=2m$, $m=1.35$ GeV and $\epsilon = 0.035$, whereas the solid line is for $\mu=2\sqrt{Q^2+4m^2}$, $m=1.65$ GeV and $\epsilon = 0.06$. From Fig.~\ref{FIGF2C3} and studies in Ref.~\cite{H1paper} it is clear that a wide range of studies can be and are being performed already at the single particle inclusive level. Preliminary conclusions \cite{H1paper,zeus} are that the data follow the shape of the NLO predictions quite well, but lie above the theory curves. The H1 collaboration \cite{H1paper} has recently shown clearly from the $d\ln\sigma/dx_D$ distribution that the charm production mechanism is indeed boson-gluon fusion, (after earlier indications from the EMC collaboration \cite{EMC}) as opposed to one where the charm quark is taken from the sea. Here $x_D=2|\vec{p}_{D^*}|/W$ in the $\gamma^* P$ c.m. frame Next we examine a few charm-anticharm correlations. \begin{figure}[htbp] \hspace*{1.5cm} \epsfig{file=fig4.ps,bbllx=0pt,bblly=70pt,bburx=575pt,bbury=800pt,% width=10cm,angle=-90} \caption[junk]{{\it Differential distributions $dF_2(x,Q^2,m^2,p_{cc})/dp_T^{cc}$ and $dF_2(x,Q^2,m^2,p_{cc})/d\Delta\phi^{cc}$ at $x=0.001$ and $Q^2 = 10$ GeV$^2$ (solid) and 100 GeV$^2$ (dashed). }} \label{FIGF2C4} \end{figure} At the experimental level such correlations are more difficult to measure since it requires the identification of both heavy quarks in the final state. However, with the expected large luminosity that both ZEUS and H1 will collect, such studies are likely to be done. As an example we show in Fig.~\ref{FIGF2C4} the $p_T$ distribution of the pair, $p_T^{cc}$, and the distribution in their azimuthal angle difference, $\Delta\phi^{cc}$ in the $\gamma^* P$ c.m. frame for a particular choice of $x$ and $Q^2$. For these figures we used the MRSA$'$ densities \cite{mrsap}. Both distributions are a measure of the recoiling gluon jet. In summary, differential distributions of deep-inelastic heavy quark production offer a rich variety of studies of the QCD production mechanism. Fruitful experimental studies, even with low statistics, have been done \cite{H1paper,zeus}, and with a large integrated luminosity we therefore fully expect many more. We finally point out that besides a LO shower Monte Carlo program \cite{aroma}, now also a NLO program is available for producing differential distributions. \section{When is Charm a Parton?} We return to the inclusive case to ask the fundamental question in the title. The question can be more accurately phrased as follows: intuitively one expects that at truly large $Q^2$ the charm quark should be described as a light quark, i.e. as a constituent parton of the proton, whereas at small $Q^2$ (of order $m^2$) the boson-gluon fusion mechanism, in which the charm quark can only be excited by a hard scattering, is the correct description. This has been demonstrated recently by H1 \cite{H1paper} and ZEUS in \cite{zeus}. In this section we examine where the transition between the two pictures occurs. At LO this issue was investigated in \cite{riol}. A picture that consistently combines both descriptions, the so-called variable flavor number scheme, is presented and worked out to LO in \cite{acot}. Here we exhibit where the transition occurs at NLO \cite{BMMSN}. In other words we will locate the onset of the large $Q^2$ asymptotic region, where the exact partonic coefficient functions of \cite{LRSNF} are dominated by large logarithms $\ln(Q^2/m^2)$. These logarithms are controlled by the renormalization group, and, when resummed, effectively constitute the charm parton density. Here we however restrict ourselves to the onset of the asymptotics. Let us be somewhat more precise. In (\ref{strfns}) we can rewrite e.g. all terms proportional to $e_H^2$ as \begin{equation} x \int_x^{zmax}\frac{dz}{z}\Big\{ \Sigma(\frac{x}{z},\mu^2) H_{i,q}(z,\frac{Q^2}{m^2},\frac{m^2}{\mu^2})+ G(\frac{x}{z},\mu^2) H_{i,g}(z,\frac{Q^2}{m^2},\frac{m^2}{\mu^2})\Big\} \end{equation} where $G(x,\mu^2)$ is the gluon density and $\Sigma(x,\mu^2) = \sum_{i=q,\bar{q}}f_{i}(x,\mu^2)$ is the singlet combination of quark densities. In the asymptotic regime one may write \begin{equation} H^{(k)}_{i,j}(z,\frac{Q^2}{m^2},\frac{m^2}{\mu^2}) = \sum_{l=0}^{k} a_{i,j}^{(k)}(z,\frac{m^2}{\mu^2}) \ln^l\frac{Q^2}{m^2}\,. \end{equation} The effort lies in determining the coefficients $a_{i,j}^{(k)}$. Similar expressions hold for the other coefficients in (\ref{strfns}). Taking the limit of the coefficients in \cite{LRSNF} is extremely complicated. Rather, a trick \cite{BMMSN} was used, exploiting the close relationship of the $\ln(Q^2/m^2)$ logarithms with collinear (mass) singularities. The ingredients are the massless two-loop coefficient functions of \cite{zn} and certain two-loop operator matrix elements. \begin{figure}[htbp] \hspace*{1.5cm} \epsfig{file=fig5.ps,bbllx=0pt,bblly=250pt,bburx=575pt,bbury=700pt,% width=12cm,angle=0} \caption[junk]{{\it Ratio of the asymptotic to exact expressions for $F_2(x,Q^2,m^2)$ for the case of charm. }} \label{FIGF2C5} \end{figure} The trick, dubbed ``inverse mass factorization'', essentially amounts to reinserting into the IR safe massless coefficient functions the collinear singularities represented by the logarithms $\ln(Q^2/m^2)$. See \cite{BMMSN} for details. There is another advantage to obtaining the asymptotic expresssions. The terms in eq.~(\ref{strfns}) proportional to $e_L^2$ have been integrated and full analytical expressions for them exist \cite{BMMSN}, but in the other terms in eq.~(\ref{strfns}) two integrals still need to be done numerically. Therefore in the large $Q^2$ region the asymptotic formula is able to give the same results much faster, as the latter formula needs no numerical integrations. In Fig.~\ref{FIGF2C5} we show the ratio of the asymptotic to exact expressions for $F_2(x,Q^2,m^2)$ for the case of charm as a function of $Q^2$ for four different $x$ values. Here the GRV \cite{GRV94} parton density set was used, for three light flavors. We see that, surprisingly, already at $Q^2$ of order 20-30 GeV$^2$ the asymptotic formula is practically identical to the exact result, indicating that at these not so large $Q^2$ values, and for the inclusive structure function, the charm quark behaves already very much like a parton. This is in apparent contradiction with the findings \cite{H1paper}, mentioned in the previous section, that the production mechanism is boson-gluon fusion, and illustrates that, interestingly, the question in the title can have a different answer for inclusive quantities than for differential distributions having multiple scales. We finally note that with the results shown in this section also the first important step is made for extending the variable flavour number scheme to NLO. \section{Conclusions} In the above we have reviewed the many interesting facets of deep-inelastic production of heavy quarks. The possibility of selecting the heavy quarks among the final state particles affords a window into the heart of the scattering process, and allows tests and measurements of some of the most fundamental aspects of perturbative QCD: the direct determination of the gluon density, many and varied studies of the heavy quark production dynamics, and insight into how and when a heavy quark becomes a parton.
proofpile-arXiv_065-403
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} The electric polarisability of the charged pion $\alpha_E$ can be inferred from the amplitude for low energy Compton scattering $ \gamma + \pi^+ \rightarrow \gamma + \pi^+ $. This amplitude cannot be measured at low energies directly, but can be determined from measurements on related processes like $\pi N \rightarrow \pi N \gamma$, $ \gamma N \rightarrow \gamma N \pi$ and $ \gamma \gamma \rightarrow \pi \pi$. The measured values for $\alpha_E$, in units of $10^{-4}\; {{\rm fm}}^3$, are $6.8 \pm 1.4 \pm 1.2$ \cite{Antipov}, $ 20 \pm 12$ \cite{Aibergenov} and $2.2 \pm 1.6$ \cite{Babusci}, respectively. Alternatively, the polarisability can be predicted theoretically by relating it to other quantities which are better known experimentally. In chiral perturbation theory, it can be shown \cite{Holstein} that the pion polarisability is given by \begin{equation} \alpha_E = \frac{ \alpha \, F_A}{m_{\pi} F_{\pi}^2}, \label{alpha} \end{equation} where $F_{\pi}$ is the pion decay constant and $F_A$ is the axial structure-dependent form factor for radiative charged pion decay $\pi \rightarrow e \nu \gamma$ \cite{Bryman}. The latter is often re-expressed in the form \begin{equation} F_A \equiv \gamma\,{F_V}\, , \end{equation} because the ratio $\gamma$ can be measured in radiative pion decay experiments more accurately than $F_A$ itself, while the corresponding vector form factor $F_V$ is determined to be \cite{PDG}\footnote{Our definitions of $F_V$ and $F_A$ differ from those used by the Particle Data Group \cite{PDG} by a factor of two.} $F_V = 0.0131 \pm 0.0003$ by using the conserved vector current (CVC) hypothesis to relate $\pi \rightarrow e \nu \gamma$ and $\pi^0 \rightarrow \gamma\gamma$ decays. $\gamma$ has been measured in three $\pi \rightarrow e \nu \gamma$ experiments, giving the values: $0.25 \pm 0.12$ \cite{lampf}; $0.52 \pm 0.06$ \cite{sin}; and $0.41 \pm 0.23$ \cite{istra}. The weighted average $ \gamma = 0.46 \pm 0.06$ can be combined with the above equations to give \begin{equation} \alpha_E = (2.80 \pm 0.36) \, 10^{-4} \; {{\rm fm}}^3 \; . \label{chiral} \end{equation} This result is often referred to as the chiral theory prediction for the pion polarisability \cite{Holstein}. However $\alpha_E$, or equivalently $F_A$, can be determined in other ways. In particular, the latter occurs in the Das-Mathur-Okubo (DMO) sum rule \cite{Das} \begin{equation} I = F_{\pi}^2 \frac{\langle r_{\pi}^2 \rangle }{3} - {F_A}\,, \label{DMO} \end{equation} where \begin{equation} I\equiv\int \frac{{\rm d}s}{s} \rho_{V-A}(s) \,, \label{i} \end{equation} with $\rho_{V-A}(s) = \rho_V(s) - \rho_A(s)$ being the difference in the spectral functions of the vector and axial-vector isovector current correlators, while $\langle r_{\pi}^2\rangle $ is the pion mean-square charge radius. Using its standard value $\langle r_{\pi}^2\rangle = 0.439 \pm 0.008 \; {{\rm fm}}^2$ \cite{Amendolia} and eqs. (\ref{alpha}), (\ref{chiral}) one gets: \begin{equation}\label{iexp} I_{DMO}=(26.6 \pm 1.0)\cdot 10^{-3} \end{equation} Alternatively, if the integral $I$ is known, eq. (\ref{DMO}) can be rewritten in the form of a prediction for the polarisability: \begin{equation} \alpha_E = \frac{\alpha}{m_{\pi}} \biggl( \frac{\langle r_{\pi}^2 \rangle }{3} - \frac{I}{F_{\pi}^2} \biggr). \label{DMOpredict} \end{equation} Recent attempts to analyse this relation have resulted in some contradiction with the chiral prediction. Lavelle et al. \cite{Lavelle} use related QCD sum rules to estimate the integral $I$ and obtain $\alpha_E = (5.60 \pm 0.50) \, 10^{-4} \; {{\rm fm}}^3$. Benmerrouche et. al. \cite{Bennerrouche} apply certain sum rule inequalities to obtain a lower bound on the polarisability (\ref{DMOpredict}) as a function of ${\langle r_{\pi}^2 \rangle }$. Their analysis also tends to prefer larger $\alpha_E$ and/or smaller ${\langle r_{\pi}^2\rangle }$ values. In the following we use available experimental data to reconstruct the hadronic spectral function $\rho_{V-A}(s)$, in order to calculate the integral \begin{equation}\label{i0} I_0(s_0) \equiv \int_{4m_{\pi}^2}^{s_0} \frac{{\rm d}s}{s} \rho_{V-A}(s) \; . \end{equation} for $s_0\simeq M_{\tau}^2$, test the saturation of the DMO sum rule (\ref{DMO}) and its compatibility with the chiral prediction (\ref{chiral}). We also test the saturation of the first Weinberg sum rule \cite{Weinberg}: \begin{eqnarray} W_1(s_0) \equiv \int_{4m_{\pi}^2}^{s_0} {\rm d}s \, \rho_{V-A}(s)\;;\;\;\;\;\;\;\;\;\;\; W_1(s_0)\;\bigl|_{s_0\to\infty} = F_{\pi}^2 \label{w1} \end{eqnarray} and use the latter to improve convergence and obtain a more accurate estimate for the integral $I$: \begin{eqnarray}\label{i1} I_1(s_0) & = & \int_{4m_{\pi}^2}^{s_0} \frac{{\rm d}s}{s} \rho_{V-A}(s) \nonumber \\ & + & {{\beta}\over{s_0}} \left[ F_{\pi}^{2} - \int_{4m_{\pi}^2}^{s_0} {\rm d}s \, \rho_{V-A}(s) \right] \end{eqnarray} Here the parameter $\beta$ is arbitrary and can be chosen to minimize the estimated error in $I_1$ \cite{markar}. Yet another way of reducing the uncertainty in our estimate of $I$ is to use the Laplace-transformed version of the DMO sum rule \cite{marg}: \begin{equation} I_2(M^2) = F_{\pi}^2 \frac{\langle r_{\pi}^2 \rangle}{3} - {F_A} \end{equation} with $M^2$ being the Borel parameter in the integral \begin{eqnarray} I_2(M^2) &\equiv& \int \frac{{\rm d}s}{s} \exp{\left( \frac{-s}{M^2} \right)} \, \rho_{V-A}(s) + \frac{F_{\pi}^2}{M^2} \nonumber \\ & - & \frac{C_6 \langle O^6\rangle}{6 M^6} - \frac{C_8 \langle O^8\rangle }{24 M^8} + \ldots \;. \label{i2} \end{eqnarray} Here $ C_6 \langle O^6\rangle $ and $ C_8 \langle O^8\rangle $ are the four-quark vacuum condensates of dimension 6 and 8, whose values could be estimated theoretically or taken from previous analyses \cite{markar,kar}. All the three integrals (\ref{i0}), (\ref{i1}) and (\ref{i2}) obviously reduce to (\ref{i}) as $s_0, M^2 \to \infty$. \section{Evaluation of the spectral densities} Recently ALEPH published a comprehensive and consistent set of $\tau$ branching fractions \cite{aleph}, where in many cases the errors are smaller than previous world averages. We have used these values to normalize the contributions of specific hadronic final states, while various available experimental data have been used to determine the shapes of these contributions. Unless stated otherwise, each shape was fitted with a single relativistic Breit-Wigner distribution with appropriately chosen threshold behaviour. \subsection{ Vector current contributions.} A recent comparative study {\cite{eidel}} of corresponding final states in $\tau$ decays and $e^+e^-$ annihilation has found no significant violation of CVC or isospin symmetry. In order to determine the shapes of the hadronic spectra, we have used mostly $\tau$ decay data, complemented by $e^+e^-$ data in some cases. {{ $\pi^-\pi^0$ :}} ${{\rm BR}}=25.30\pm0.20\%$ \cite{aleph}, and the $s$-dependence was described by the three interfering resonances $\rho(770)$, $\rho(1450)$ and $\rho(1700)$, with the parameters taken from \cite{PDG} and \cite{pi2}. {{ $3\pi^{\pm}\pi^0$ :}} ${{\rm BR}}=4.50\pm0.12\%$, including $\pi^-\omega$ final state \cite{aleph}. The shape was determined by fitting the spectrum measured by ARGUS \cite{pi4a}. {{ $\pi^{-}3\pi^0$ :}} ${{\rm BR}}=1.17\pm0.14\%$ \cite{aleph}. The $s$-dependence is related to that of the reaction $e^+e^-\to 2\pi^+2\pi^-$. We have fitted the latter measured by OLYA and DM2 \cite{pi4b}. {{ $6\pi$ :}} various charge contributions give the overall ${{\rm BR}}=0.13\pm0.06\%$ \cite{aleph}, fairly close to CVC expectations \cite{eidel}. {{ $\pi^-\pi^0\eta$ :}} ${{\rm BR}}=0.17\pm0.03\%$ \cite{pi0}. The $s$-dependence was determined by fitting the distribution measured by CLEO \cite{pi0}. {{ $K^-K^0$ :}} ${{\rm BR}}=0.26\pm0.09\%$ \cite{aleph}. Again, the fit of the CLEO measurement \cite{cleok} was performed. \subsection{ Axial current contributions.} The final states with odd number of pions contribute to the axial-vector current. Here, $\tau$ decay is the only source of precise information. {{ $\pi^-$ :}} ${{\rm BR}}=11.06\pm0.18\%$ \cite{aleph}. The single pion contribution has a trivial $s$-dependence and hence is explicitly taken into account in theoretical formulae. The quoted branching ratio corresponds to $F_{\pi}=93.2$ MeV. {{ $3\pi^{\pm}$ and $\pi^-2\pi^0$ :}} ${{\rm BR}}=8.90\pm0.20\%$ and ${{\rm BR}}=9.21\pm0.17\%$, respectively \cite{aleph}. Theoretical models \cite{pi3th} assume that these two modes are identical in both shape and normalization. The $s$-dependence has been analyzed in \cite{opal}, where the parameters of two theoretical models describing this decay have been determined. We have used the average of these two distributions, with their difference taken as an estimate of the shape uncertainty. {{ $3\pi^{\pm}2\pi^0$ :}} ${{\rm BR}}=0.50\pm0.09\%$, including $\pi^-\pi^0\omega$ final state \cite{aleph}. The shape was fitted using the CLEO measurement \cite{pi5a}. {{ $5\pi^{\pm}$ and $\pi^-4\pi^0$ :}} ${{\rm BR}}=0.08\pm0.02\%$ and $BR=0.11\pm0.10\%$, respectively \cite{aleph}. We have assumed that these two terms have the same $s$-dependence measured in \cite{pi5b}. \subsection{ $K{\overline K} \pi$ modes.} $K{\overline K} \pi$ modes can contribute to both vector and axial-vector currents, and various theoretical models cover the widest possible range of predictions \cite{kkpith}. According to \cite{aleph}, all three $K{\overline K} \pi$ modes (${\overline K}^0K^0\pi^-, K^-K^0\pi^0$ and $K^-K^+\pi^-$) add up to BR$({\overline K} K \pi)=0.56\pm0.18\%$, in agreement with other measurements (see \cite{cleok}). The measured $s$-dependence suggests that these final states are dominated by $K^*K$ decays \cite{cleok}. We have fitted the latter, taking into account the fact that due to parity constraints, vector and axial-vector $K^*K$ terms have different threshold behaviour. A parameter $\xi$ was defined as the portion of $K{\overline K} \pi$ final state with axial-vector quantum numbers, so that \begin{eqnarray} {\rm BR}({\overline K} K\pi)_{V}&=&(1-\xi)\;{\rm BR}({\overline K}K\pi) \nonumber\\ {\rm BR}({\overline K} K\pi)_{A}&=&\xi\;{\rm BR}({\overline K} K\pi)\,. \end{eqnarray} \section{Results and conclusions} \begin{figure}[htb] \begin{center} \epsfig{file=fig1.eps,width=10cm,clip=} \end{center} \vspace{1cm} \caption{\tenrm Difference of vector and axial-vector hadronic spectral densities. In figs.1-5: the three curves correspond to $\xi=0$, $0.5$ and 1 from top to bottom; the errors originating from the shape variation and those coming from the errors in the branching fractions are roughly equal and have been added in quadrature to form the error bars, shown only for $\xi=0.5$.} \label{fig1} \end{figure} \begin{figure}[p] \begin{center} \epsfig{file=fig2.eps,width=10cm,clip=} \end{center} \vspace{1cm} \caption{\tenrm Saturation of the DMO sum rule integral (8). The thick dashed line is the chiral prediction for the asymptotic value (6) and the dotted lines show its errors.} \label{fig2} \begin{center} \epsfig{file=fig3.eps,width=10cm,clip=} \end{center} \vspace{1cm} \caption{\tenrm Saturation of the first Weinberg sum rule (9). The dashed line shows the expected asymptotic value $F_{\pi}^2$.} \label{fig3} \end{figure} The resulting spectral function is shown in Fig.1. The results of its integration according to (\ref{i0}) are presented in fig.2 as a function of the upper bound $s_0$. One can see that as $s_0$ increases, $I_0$ converges towards an asymptotic value which we estimate to be{\footnote{In the following, the first error corresponds to the quadratic sum of the errors in the branching ratios and the assumed shapes, while the second one arises from to the variation of $\xi$ in the interval $0.5\pm 0.5$.} \begin{equation}\label{i0m} I_0 \equiv I_0(\infty) = ( 27.5 \pm 1.4 \pm 1.2 ) \cdot 10^{-3}, \end{equation} in good agreement with the chiral value (\ref{iexp}). The saturation of the Weinberg sum rule (\ref{w1}) is shown in fig.3. One sees that the expected value $F_{\pi}^2$ is well within the errors, and $\xi \simeq 0.25\div0.3$ seems to be preferred. No significant deviation from this sum rule is expected theoretically \cite{Floratos}, so we use (\ref{i1}) to calculate our second estimate of the integral $I$. The results of this integration are presented in fig.4, with the asymptotic value \begin{equation}\label{i1m} I_1 \equiv I_1(\infty) = ( 27.0 \pm 0.5 \pm 0.1 ) \cdot 10^{-3}, \end{equation} corresponding to $\beta\approx 1.18$. One sees that the convergence has improved, the errors are indeed much smaller, and the $\xi$-dependence is very weak. \begin{figure}[p] \begin{center} \epsfig{file=fig4.eps,width=10cm,clip=} \end{center} \vspace{1cm} \caption{\tenrm Saturation of the modified DMO sum rule integral (10). The chiral prediction is also shown as in fig.2.} \label{fig4} \begin{center} \epsfig{file=fig5.eps,width=10cm,clip=} \end{center} \vspace{1cm} \caption{\tenrm The Laplace-transformed sum rule (12) as a function of the Borel parameter $M^2$, compared to the chiral prediction.} \label{fig5} \end{figure} Now we use (\ref{i}) to obtain our third estimate of the spectral integral. The integration results are plotted against the Borel parameter $M^2$ in fig.5, assuming standard values for dimension 6 and 8 condensates. The results are independent of $M^2$ for $M^2 > 1 \, GeV^2$, indicating that higher order terms are negligible in this region, and giving \begin{equation}\label{i2m} I_2\equiv I_2(\infty) = ( 27.2 \pm 0.4 \pm 0.2 \pm 0.3) \, 10^{-3} \; , \end{equation} where the last error reflects the sensitivity of (\ref{i}) to the variation of the condensate values. One sees that these three numbers (\ref{i0m}) -- (\ref{i2m}) are in good agreement with each other and with the chiral prediction (\ref{iexp}). Substitution of our most precise result (\ref{i1m}) into (\ref{DMOpredict}) yields for the standard value of the pion charge radius quoted above: \begin{equation}\label{alem} \alpha_E = ( 2.64 \pm 0.36 ) \, 10^{-4} \; {{\rm fm}}^3, \end{equation} in good agreement with (\ref{chiral}) and the smallest of the measured values, \cite{Babusci}. Note that by substituting a larger value $\langle r_{\pi}^2\rangle = 0.463 \pm 0.006 \; {{\rm fm}}^2$ \cite{gesh}, one obtains $\alpha_E = (3.44 \pm 0.30) \, 10^{-4} \; {{\rm fm}}^3$, about two standard deviations higher than (\ref{chiral}). In conclusion, we have used recent precise data to reconstruct the difference in vector and axial-vector hadronic spectral densities and to study the saturation of Das-Mathur-Okubo and the first Weinberg sum rules. Two methods of improving convergence and decreasing the errors have been used. Within the present level of accuracy, we have found perfect consistence between $\tau$ decay data, chiral and QCD sum rules, the standard value of $\langle r_{\pi}^2\rangle$, the average value of $\gamma$ and the chiral prediction for $\alpha_E$. Helpful discussions and correspondence with R. Alemany, R. Barlow, M.Lavelle and P. Poffenberger are gratefully acknowledged. \newpage
proofpile-arXiv_065-404
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Dynamics and time orientation} We introduce a treatment of dynamics which may be readily generalized for a subsequent application to cosmology. \subsection{Predynamical and physical time} Dynamics in general is time dependence of a state of a physical system. A definition of the state may involve the direction of time, which is not given a priori. Therefore we introduce a predynamical time, or pretime, for short, $\tau$ as a point of the oriented real axis $T$. For the direction of physical time, $t$, there are two possibilities: $t=\tau$ and $t=-\tau$. Definitions of the state and dynamics should be given in terms of the pretime, and the problem of physical time orientation is to be solved on the basis of dynamics. \subsection{Dynamical process} We start with the notion of a dynamical process, which plays a central role in dynamics. Let $\Omega$ be a set of pure states $\omega$, $\Delta$ be a connected subset of $T$, i.e., an interval: \begin{equation} \Delta=(\tau_{1},\tau_{2}),[\tau_{1},\tau_{2}), (\tau_{1},\tau_{2}],[\tau_{1},\tau_{2}],\quad -\infty\leq \tau_{1}<\tau_{2}\leq\infty. \label{1.2.1} \end{equation} A dynamical process ${\cal P}_{\Delta}$ on ${\Delta}$ is a function from $\Delta$ to $\Omega$:\begin{equation} {\cal P}_{\Delta}:\Delta\rightarrow \Omega,\quad \Delta\ni\tau \mapsto {\cal P}_{\Delta}(\tau)=\omega_{\tau}\in\Omega. \label{1.2.2} \end{equation} In fact, it would suffice for ${\cal P}_{\Delta}(\tau)$ to be defined almost everywhere on $\Delta$. A restriction and extension of a process are defined as those of a function with regard to the fact that the domain of the process is connected. A left (right) prolongation of a process ${\cal P}_{\Delta}$ to $\Delta'$, $\Delta'\ni\tau'<\tau\in\Delta$ ($\Delta\ni\tau <\tau'\in\Delta'$), is a process ${\cal P}_{\Delta'}$, such that there exists a process ${\cal P}_{\Delta\cup\Delta'}$ with the restrictions ${\cal P}_{\Delta}$ and ${\cal P}_{\Delta'}$. \subsection{Dynamics} Dynamics on $\Delta$, ${\cal D}_{\Delta}$, is a family of processes on $\Delta$: \begin{equation} {\cal D}_{\Delta}=\left\{ {\cal P}_{\Delta} \right\}. \label{1.3.1} \end{equation} A restriction and extension of a dynamics boil down to those of corresponding processes. \subsection{Deterministic process and deterministic dynamics} A deterministic process ${\cal P}_{\Delta}$ is defined as follows: For every restriction of ${\cal P}_{\Delta}$ the only extension to $\Delta$ is ${\cal P}_{\Delta}$ itself. A deterministic dynamics is a family of deterministic processes. \subsection{Indeterministic point, process, and dynamics} An interior isolated indeterministic point $\tau\in{\rm int}\:\Delta$ of a process ${\cal P}_{\Delta}$ is defined as follows: There exists $\theta>0$, such that (i) $(\tau-\theta,\tau+\theta)\subset\Delta$; (ii) left prolongations of ${\cal P}_{\Delta}|_{[\tau, \tau+\theta) }$ to ($\tau-\theta,\tau$) and right ones of ${\cal P}_{\Delta}|_ {(\tau-\theta,\tau]}$ to ($\tau,\tau+\theta$) are deterministic processes; (iii) cardinal numbers ${\rm card^{left}}$ and ${\rm card^{right}}$ of sets of those prolongations meet the condition ${\rm card^{left}+card^{right}}>2$. We assume that there are only isolated indeterministic points. An indeterministic process is that with indeterministic points. An indeterministic dynamics is one with indeterministic processes. \subsection{Orientable dynamics and time orientation: The future and the past} Let $\tau\in{\rm int}\:\Delta$ be an indeterministic point of a process ${\cal P}_{\Delta}$. We introduce five dynamics related to the point as follows: (i) ${\cal D}_{(\tau-\theta,\tau)}$, ${\cal D}_{(\tau, \tau+\theta)}$ are deterministic, ${\cal D}_{(\tau-\theta,\tau)}\ni {\cal P}_{\Delta}|_{ (\tau-\theta,\tau)}$, ${\cal D}_{(\tau,\tau+\theta)}\ni {\cal P}_{\Delta}|_{(\tau,\tau+\theta)}$; (ii) ${\cal D}_{(\tau-\theta,\tau+\theta)}=\{{\cal P} _{(\tau-\theta,\tau+\theta)}^{\alpha},\alpha\in {\cal A}\}$, ${\cal D}_{(\tau-\theta,\tau+\theta)}|_{(\tau-\theta,\tau)}= {\cal D}_{(\tau-\theta,\tau)}$, ${\cal D}_{(\tau-\theta,\tau+ \theta)}|_{(\tau,\tau+\theta)}={\cal D}_{(\tau,\tau+\theta)}$; (iii) a graph where points are elements of ${\cal D}_{(\tau- \theta,\tau)}$ and ${\cal D}_{(\tau,\tau+\theta)}$ and lines connecting related points are elements of ${\cal D} _{(\tau-\theta,\tau+\theta)}$ is connected and complete, i.e., involves all processes associated with the indeterministic point; (iv) ${\cal D}_{(\tau-\theta,\tau+\theta)}^{{\rm right}\; \alpha}= \{{\cal P}^{\alpha'}\in{\cal D}_{(\tau-\theta,\tau+\theta)}: {\cal P}^{\alpha'}|_{(\tau-\theta,\tau)}={\cal P}^{\alpha}| _{(\tau-\theta,\tau)}\}$, ${\cal D}_{(\tau-\theta,\tau+\theta)}^{{\rm left}\;\alpha}= \{{\cal P}^{\alpha'}\in{\cal D}_{(\tau-\theta,\tau+\theta)}: {\cal P}^{\alpha'}|_{(\tau,\tau+\theta)}={\cal P}^{\alpha}| _{(\tau,\tau+\theta)}\}$, ${\rm card^{right\;\alpha},\; card^{left\;\alpha}}$ being corresponding cardinal numbers. An indeterministic point is symmetric (asymmetric) if ${\rm card^{right\;\alpha}=(\ne)card^{left\;\alpha}}$. A symmetric dynamics is that with symmetric indeterministic points only. An indeterministic dynamics is orientable if for all indeterministic points either \begin{equation} {\rm card^{right\;\alpha}}> {\rm card^{left\;\alpha}}\quad{\rm and}\quad {\rm card}\:D_{(\tau,\tau+\theta)}> {\rm card}\:D_{(\tau-\theta,\tau)} \label{1.6.1} \end{equation} or \begin{equation} {\rm card^{right\;\alpha}<card^{left\;\alpha}}\quad {\rm and}\quad {\rm card}\:D_{(\tau,\tau+\theta)}< {\rm card}\:D_{(\tau-\theta,\tau)}. \label{1.6.2} \end{equation} An orientable dynamics is oriented as follows: The future corresponds to the greater of ${\rm card^{right\;\alpha}}$, ${\rm card}\:D_{(\tau,\tau+\theta)}$ and ${\rm card^{left\;\alpha}}$, ${\rm card}\:D_{(\tau-\theta,\tau)}$, i.e., \begin{equation} {\rm card^{future\;\alpha}>card^{past\;\alpha}},\quad {\rm card}\:D_{\rm future}>{\rm card}\:D_{\rm past}; \label{1.6.3} \end{equation} so that the physical time is \begin{equation} t=+(-)\tau\quad {\rm for}\;{\rm card^{right\;\alpha}>(<) \;card^{left\;\alpha}},\; {\rm card}\:D_{(\tau,\tau+\theta)}>(<)\; {\rm card}\:D_{(\tau-\theta,\tau)}. \label{1.6.4} \end{equation} This defines time orientation, or the arrow of time. \subsection{Nonpredeterminability and the question of reconstructibility} For an oriented dynamics we have \begin{equation} {\rm card^{future\;\alpha}+card^{past\;\alpha}}>2, \label{1.7.1} \end{equation} \begin{equation} {\rm card^{future\;\alpha}>card^{past\;\alpha}}\geq 1, \label{1.7.2} \end{equation} so that \begin{equation} {\rm card^{future\;\alpha}}>1. \label{1.7.3} \end{equation} This implies that the future is not predeterminate. If \begin{equation} {\rm card^{past\;\alpha}}=1, \label{1.7.4} \end{equation} the past is reconstructible. In any case, the inequality (\ref{1.6.1}) implies that the reconstructibility of the past is greater than the predictability of the future. This feature is inherent in an oriented dynamics. The phenomenon of memory should be related to dynamics orientation. \subsection{Probabilistic dynamics} Let $\tau$ be an indeterministic point of a process ${\cal P}_{\Delta}$. Time evolution implies transitions from one of the sets ${\cal D}_{(\tau-\theta,\tau)}$, ${\cal D}_{(\tau, \tau+\theta)}$ to the other: from ${\cal D}^{\rm initial}$ to ${\cal D} ^{\rm final}$. We assume that for their cardinal numbers \begin{equation} {\rm card^{final}\geq card^{initial}} \label{1.8.1} \end{equation} holds. Let there exist $i\to f$ transition probabilities, or conditional probabilities $w(f/i)$, where $i$ and $f$ are indexes of elements of ${\cal D}^{\rm initial}$ and ${\cal D}^{\rm final}$ respectively. The probabilities meet the equation \begin{equation} \sum_{f}w(f/i)=1. \label{1.8.2} \end{equation} Taking into account the relations \begin{equation} {\rm card^{initial}}=\sum_{i}1=\sum_{i}\sum_{f}w(f/i)= \sum_{f}\sum_{i}w(f/i)\leq \sum_{f}1={\rm card^{final}}, \label{1.8.3} \end{equation} we put \begin{equation} \sum_{i}w(f/i)\leq 1. \label{1.8.4} \end{equation} By Bayes formula, the a posteriori probability is \begin{equation} w(i/f)=\frac{w(i)w(f/i)}{\sum_{i'}w(i')w(f/i')}. \label{1.8.5} \end{equation} We put for the a priori probability \begin{equation} w(i)={\rm const}, \label{1.8.6} \end{equation} then \begin{equation} w(i/f)=\frac{w(f/i)}{\sum_{i'}w(f/i')}\geq w(f/i). \label{1.8.7} \end{equation} Thus \begin{equation} w(f/i)\leq w(i/f)\leq 1. \label{1.8.8} \end{equation} For a symmetric dynamics, \begin{equation} {\rm card^{final}=card^{initial}},\quad \sum_{i}w(f/i)=1, \label{1.8.9} \end{equation} so that \begin{equation} w(i/f)=w(f/i). \label{1.8.10} \end{equation} Specifically, the increase of entropy is, on the average, the same for the future and for the past: \begin{equation} -\sum_{f}w(f/i)\ln w(f/i)\approx-\sum_{i}w(i/f)\ln w(i/f). \label{1.8.11} \end{equation} \subsection{Irreversibility and orientation} It should be particularly emphasized that irreversibility does not imply dynamics orientability and, by the same token, time orientation. Indeed, a reversible dynamics is defined as follows. Let ${\cal P}_{\Delta}$ be a process with a symmetric domain, i.e., $\Delta=(\tau_{1},\tau_{2})$ or $[\tau_{1},\tau_{2}]$. The inverse process, ${\cal P}_{\Delta}^{\rm inv}$, is defined by \begin{equation} {\cal P}_{\Delta}^{\rm inv}(\tau)={\cal P}_{\Delta}(\tau_{1}+ \tau_{2}-\tau), \quad \tau\in\Delta. \label{1.9.1} \end{equation} Let $S$ be a transformation of $\Omega$, $S:\Omega\to\Omega$. The transformed process, $S{\cal P}_{\Delta}$, is defined by \begin{equation} S{\cal P}_{\Delta}(\tau)=S({\cal P}_{\Delta}(\tau)),\quad \tau\in\Delta. \label{1.9.2} \end{equation} A dynamics ${\cal D}_{\Delta'}$ is reversible if there exists a bijection $S:\Omega\to\Omega$, such that \begin{equation} S\;{\rm is\;an\;involution}\;(S^{2}\;{\rm is\; identity) \; and}\; {\cal P} _{\Delta}\in {\cal D}_{\Delta} \Rightarrow {\cal P}_{\Delta}^{\rm rev}\equiv S{\cal P}_ {\Delta}^{\rm inv} \in {\cal D}_{\Delta}\; {\rm for\; all}\; \Delta\subset \Delta' \label{1.9.3} \end{equation} (rev stands for reverse). The nonexistence of $S$ does not imply the orientability of ${\cal D}_{\Delta'}$. Here is an example. Let a dynamical equation be of the form \begin{equation} \frac{d^{2}x}{d\tau^{2}}=-\alpha\frac{dx}{d\tau}. \label{1.9.4} \end{equation} All dynamical processes ${\cal P}_{(-\infty,\infty)}$ are given by \begin{equation} {\cal P}_{(-\infty,\infty)}(\tau)=\omega_{\tau}=\left( x(\tau), \frac{dx(\tau)}{d\tau} \right), \label{1.9.5} \end{equation} \begin{equation} x(\tau)=c_{1}+c_{2}e^{-\alpha\tau}; \label{1.9.6} \end{equation} they are deterministic. The dynamics ${\cal D}_{(-\infty,\infty)}$ is irreversible but deterministic and, therefore, not orientable. On the other hand, a dynamics with asymmetric indeterministic points is irreversible---in view of inequality ${\rm card ^{right\:\alpha}\ne card^{left\:\alpha}}$. Specifically, an orientable dynamics is irreversible. \section{Dynamics of standard indeterministic quantum theory} Let us consider the dynamics of standard, or orthodox indeterministic quantum theory from the standpoint developed in the previous section. \subsection{Standard dynamical process} In standard quantum theory, indeterminism originates from quantum jumps. A standard indeterministic dynamical process ${\cal P}_{(-\infty,\infty)}$ may be described as follows. Let $\tau_{k},\;k\in K=\{0,\pm 1,\pm 2,...\}$, be indeterministic points, i.e., points of jumps. The process is denoted by \begin{equation} {\cal P}_{(-\infty,\infty)}^{\{j_{k},k\in K\}},\;j_{k}\in J=\{1,2,..., j_{\rm max}\},\;j_{\rm max}\leq \infty. \label{2.1.1} \end{equation} The definition of this process reduces to that of its restrictions to the intervals \begin{equation} \Delta_{k}=(\tau_{k},\tau_{k+1}),\quad k\in K, \label{2.1.2} \end{equation} \begin{equation} {\cal P}_{\Delta_{k}}^{j_{k}}\equiv {\cal P}_{(-\infty, \infty)}^{\{j _{k},k\in K\}}|_{\Delta_{k}}. \label{2.1.3} \end{equation} The process (\ref{2.1.3}) is defined as follows: \begin{equation} {\cal P}_{\Delta_{k}}^{j_{k}}(\tau)=\omega^{j_{k}}_{\tau}= (\Psi^{j_{k}} (\tau),\cdot\Psi^{j_{k}}(\tau)),\quad \tau\in\Delta_{k}, \label{2.1.4} \end{equation} where $\Psi^{j_{k}}$ is a state vector, \begin{equation} \Psi^{j_{k}}(\tau)=U(\tau,\tau_{k})\Psi_{j_{k}}, \label{2.1.5} \end{equation} \begin{equation} A_{k}\Psi_{j_{k}}=a_{j_{k}}\Psi_{j_{k}}, \label{2.1.6} \end{equation} where the $A_{k}$ is an observable, and $U$ is a unitary operator of time evolution. This description seemingly fixes the time orientation, namely, in view of eq.(\ref{2.1.5}), \begin{equation} t=\tau. \label{2.1.7} \end{equation} But there is another possibility for describing the process considered. \subsection{Reverse description} In place of eqs.(\ref{2.1.5}),(\ref{2.1.6}), we may put \begin{equation} \Psi^{j_{k}}(\tau)=U(\tau,\tau_{k+1})\Psi_{j_{k+1}}^{\rm rev}, \label{2.2.1} \end{equation} \begin{equation} A_{k+1}^{\rm rev}\Psi_{j_{k+1}}^{\rm rev}=a_{j_{k+1}} ^{\rm rev}\Psi_{j_{k+1}}^{\rm rev}, \label{2.2.2} \end{equation} where \begin{equation} \Psi_{j_{k+1}}^{\rm rev}=U(\tau_{k+1},\tau_{k})\Psi_{j_{k}}, \label{2.2.3} \end{equation} \begin{equation} A_{k+1}^{\rm rev}=U(\tau_{k+1},\tau_{k})A_{k}U(\tau_{k}, \tau_{k+1}), \label{2.2.4} \end{equation} \begin{equation} a_{j_{k+1}}^{\rm rev}=a_{j_{k}}. \label{2.2.5} \end{equation} This description implies, in view of eq.(\ref{2.2.1}), the time orientation \begin{equation} t=-\tau. \label{2.2.6} \end{equation} The two descriptions are completely equivalent physically. \subsection{Standard quantum dynamics} We have for an indeterministic point $\tau_{k}$ \begin{equation} {\cal D}_{k}^{\rm left}\equiv{\cal D} _{(\tau_{k}-\theta,\tau_{k})}= \{{\cal P}_{\Delta_{k-1}}^{j_{k-1}}|_{(\tau_{k}-\theta, \tau_{k})} ,j_{k-1}\in J\}, {\cal D}_{k}^{\rm right}\equiv{\cal D} _{(\tau_{k},\tau_{k}+\theta)}= \{{\cal P}_{\Delta_{k}}^{j_{k}}|_{(\tau_{k},\tau_{k}+\theta)}, j_{k}\in J\}, \label{2.3.1} \end{equation} \begin{equation} {\rm card}\:{\cal D}_{k}^{\rm left}={\rm card}\:{\cal D}_{k} ^{\rm right}= {\rm card}\:J. \label{2.3.2} \end{equation} Thus standard quantum dynamics is not orientable. \subsection{Standard probabilistic quantum dynamics} We have for the time orientation $t=\tau$ \begin{equation} w(j_{k+1}/j_{k})=w_{j_{k+1}\gets j_{k}}=|(\Psi^{j_{k+1}} (\tau_{k+1}+0),\Psi^{j_{k}}(\tau_{k+1}-0))|^{2}= |(\Psi_{j_{k+1}},U(\tau_{k+1},\tau_{k})\Psi_{j_{k}})|^{2}, \label{2.4.1} \end{equation} \begin{equation} w_{j_{k+m}\gets j_{k+m-1}\gets\ldots\gets j_{k+1}\gets j_{k}} =w(j_{k+m}/j_{k+m-1})\cdots w(j_{k+1}/j_{k}); \label{2.4.2} \end{equation} for the time orientation $t=-\tau$ \begin{equation} w(j_{k}/j_{k+1})=w_{j_{k}\gets j_{k+1}}= |(\Psi^{j_{k}}(\tau_{k+1}-0),\Psi^{j_{k+1}}(\tau_{k+1}+0))| ^{2}=w(j_{k+1}/j_{k}), \label{2.4.3} \end{equation} \begin{equation} w_{j_{k}\gets\ldots\gets j_{k+m}}=w_{j_{k+m}\gets\ldots\gets j _{k}}. \label{2.4.4} \end{equation} The probabilities satisfy the equations \begin{equation} \sum_{j_{k+1}}w(j_{k+1}/j_{k})=\sum_{j_{k}} w(j_{k+1}/j_{k})=1. \label{2.4.5} \end{equation} We obtain for $t=\tau$ by Bayes formula, under the condition $w(j_{k})={\rm const}$, \begin{equation} w(j_{k}/j_{k+1})=w(j_{k+1}/j_{k}), \label{2.4.6} \end{equation} which coincides with eq.(\ref{2.4.3}). \subsection{Nonorientability of standard indeterministic quantum dynamics} Summing up the results of this section, we conclude that the dynamics of standard indeterministic quantum theory is nonorientable and, by the same token, does not fix the orientation of physical time. \section{Dynamics of indeterministic quantum gravity} As in standard quantum theory, in indeterministic quantum gravity indeterminism originates from quantum jumps. But the origin of the jumps in the latter theory differs radically from that in the former one. A quantum jump is the reduction of a state vector to one of its components. In standard quantum theory, the cause of the jump is coherence breaking between the components. In indeterministic quantum gravity, the cause is energy difference between the components, the difference occurring at a crossing of energy levels. According to the paper [2], a jump occurs at the tangency of two levels. But level tangency imposes too severe constraints on the occurrence of the jump. Here we introduce a scheme in which the jump occurs at a simple crossing of two levels. \subsection{Level crossing} Let $\tau=0$ be the point of a crossing of levels $l=1,2$; $P_{1\tau},P_ {2\tau}$ be the projectors for the corresponding states in a neighborhood of the point: \begin{equation} P_{l\tau}\leftrightarrow \omega_{ml\tau}= (\Psi_{l\tau},\cdot\Psi_{l\tau}) \label{3.1.1} \end{equation} ($m$ stands for matter), and \begin{equation} P_{\tau}=P_{1\tau}+P_{2\tau}. \label{3.1.2} \end{equation} The part of the Hamiltonian $H_{\tau}$ related to the two levels is a projected Hamiltonian \begin{equation} H_{\tau}^{\rm proj}=P_{\tau}H_{\tau}P_{\tau}= \epsilon_{1\tau}P_{1\tau}+\epsilon_{2\tau}P_ {2\tau} \label{3.1.3} \end{equation} $(H_{\tau}^{\rm proj}\;{\rm is}\;\tilde H_{t}\; {\rm in}\;[2])$. The metric tensor is \begin{equation} g=d\tau\otimes d\tau-h_{\tau} \label{3.1.4} \end{equation} $(h_{\tau}\;{\rm is}\;\tilde g_{t}\;{\rm in}\;[2])$. We have \begin{equation} H_{\tau}^{\rm proj}=H^{\rm proj}[h_{\tau},\dot h_{\tau}], \label{3.1.5} \end{equation} where dot denotes the derivative with respect to the pretime $\tau$, \begin{equation} H_{0}^{\rm proj}=\epsilon_{0}P_{0}=H^{\rm proj} [h_{0},\dot h_{0}],\quad \epsilon_{0}=\epsilon_{10}= \epsilon_{20}. \label{3.1.6} \end{equation} \subsection{Creation projector and creation state} We have in the first order in $\tau$ \begin{equation} H_{\tau}^{\rm proj}=H_{0}^{\rm proj}+\dot H_{0}^{\rm proj}\tau =H_{0}^{\rm proj}+\dot H^{\rm proj}[h_{0},\dot h_{0},\ddot h _{0}]\tau. \label{3.2.1} \end{equation} Furthermore, \begin{equation} \ddot h_{0}=\ddot h[h_{0},\dot h_{0},P^{\rm creat}], \label{3.2.2} \end{equation} where $P^{\rm creat}$ is a one-dimensional projector which creates $\ddot h_{0}$ and, by the same token, the Hamiltonian $H_{\tau}^{\rm proj}$ eq.(\ref{3.2.1}). This creation projector satisfies \begin{equation} P^{\rm creat}P_{0}=P^{\rm creat} \label{3.2.3} \end{equation} and corresponds to a creation state $\omega_{m}^{\rm creat}$ belonging to a state subspace determined by $P_{0}$. For the sake of brevity, we write \begin{equation} H_{\tau}^{\rm proj}=H_{0}^{\rm proj}+v\tau, \label{3.2.4} \end{equation} \begin{equation} v=\dot H^{\rm proj}[h_{0},\dot h_{0},\ddot h[h_{0}, \dot h_{0},P^{\rm creat}]]. \label{3.2.5} \end{equation} \subsection{Diagonal Hamiltonian} The diagonalization of the Hamiltonian $H_{\tau}^{\rm proj}$ eq.(\ref{3.2.4}) gives \begin{equation} H_{\tau}^{\rm proj}=(\epsilon_{0}+\epsilon_{\tau}^{+}) P^{+} +(\epsilon_{0}+\epsilon_{\tau}^{-})P^{-}, \label{3.3.1} \end{equation} where \begin{equation} \epsilon_{\tau}^{\pm}=\tau\frac{v_{11}+ v_{22}}{2}\pm |\tau|\sqrt{\frac{(v_{11}-v_{22})^{2}}{4} +|v_{12}|^{2}}, \label{3.3.2} \end{equation} \begin{equation} P^{\pm}\leftrightarrow\omega_{m}^{\pm}=(\Psi^{\pm},\cdot \Psi^{\pm}), \label{3.3.3} \end{equation} \begin{equation} \Psi^{+}=e^{{\rm i}\beta}\cos\vartheta\: \Psi_{1}+\sin\vartheta \:\Psi_{2},\quad\Psi^{-}=-e^{{\rm i}\beta}\sin\vartheta\: \Psi_{1} +\cos\vartheta\:\Psi_{2}, \label{3.3.4} \end{equation} \begin{equation} e^{{\rm i}\beta}=\frac{\tau}{|\tau|}\frac{v_{12}}{|v_{12}|}, \label{3.3.5} \end{equation} \begin{equation} \tan\vartheta=\frac{|v_{12}|}{(\tau/|\tau|)(v_{11}-v_{22})/2 +\sqrt{(v_{11}-v_{22})^{2}/4+|v_{12}|^{2}}}, \label{3.3.6} \end{equation} \begin{equation} \cot\vartheta=\frac{|v_{12}|}{(-\tau/|\tau|)(v_{11}-v_{22}) /2 +\sqrt{(v_{11}-v_{22})^{2}/4+|v_{12}|^{2}}}, \label{3.3.7} \end{equation} \begin{equation} v_{ll'}=(\Psi_{l},v\Psi_{l'}), \label{3.3.8} \end{equation} and $\{\Psi_{1},\Psi_{2}\}$ is a basis in the two-dimensional Hilbert subspace ${\cal H}_{0}^{(2)}$ determined by $P_{0}$. We have \begin{equation} \epsilon_{-\tau}^{\pm}=-\epsilon_{\tau}^{\mp}, \label{3.3.9} \end{equation} \begin{equation} \tau\to -\tau\Rightarrow e^{{\rm i}\beta}\to-e^{{\rm i}\beta}, \tan\vartheta\leftrightarrow \cot\vartheta,\sin\vartheta \leftrightarrow\cos\vartheta,\Psi^{+}\leftrightarrow\Psi^{-}, P^{+}\leftrightarrow P^{-}. \label{3.3.10} \end{equation} \subsection{Germ projector, germ state, and germ process} A germ projector $P^{\rm germ}$ is one of the two projectors $P^{\pm}$ eq.(\ref{3.3.3}); it gives rise to a germ process ${\cal P}^{\rm germ}$---a process in a proximity of the point $\tau=0$; ${\cal P}_{(0,\theta)}^{\rm germ\;right}$ and ${\cal P}_{(-\theta,0)}^{\rm germ\;left}$ are defined by (i) ${\cal P}^{\rm germ}$ is deterministic; (ii) $\lim_{\tau\to+0}{\cal P}_{(0,\theta)}^{\rm germ\;right} (\tau)=\omega_{m}^{\rm germ\;right}\leftrightarrow P ^{\rm germ\;right}$; (iii) $\lim_{\tau\to-0}{\cal P}_{(-\theta,0)}^{\rm germ\; left} (\tau)=\omega_{m}^{\rm germ\;left}\leftrightarrow P ^{\rm germ\;left}$. For $\Psi\in{\cal H}_{0}^{(2)}$ we put \begin{equation} \Psi=e^{{\rm i}\alpha}\cos\varphi\:\Psi_{1}+\sin \varphi\:\Psi_{2}, \label{3.4.1} \end{equation} so that \begin{equation} P^{\rm creat}\leftrightarrow(\alpha,\varphi). \label{3.4.2} \end{equation} The operator $v$ eq.(\ref{3.2.5}) is a function of $(\alpha,\varphi)$, so that $\beta,\theta$ eqs.(\ref{3.3.5}),(\ref{3.3.6}),(\ref{3.3.7}) are such functions as well, \begin{equation} (\alpha,\varphi)\to(\beta,\theta). \label{3.4.3} \end{equation} We assume that there exist the inverse functions, \begin{equation} (\beta,\theta)\to(\alpha,\varphi), \label{3.4.4} \end{equation} so that there exists a bijection \begin{equation} (\alpha,\varphi)\leftrightarrow(\beta,\theta). \label{3.4.5} \end{equation} As \begin{equation} P^{+}+P^{-}=P_{0}, \label{3.4.6} \end{equation} so that \begin{equation} P^{+}\leftrightarrow P^{-} \label{3.4.7} \end{equation} and \begin{equation} P^{\rm germ}\to\{P^{+},P^{-}\}\leftrightarrow (\beta,\theta), \label{3.4.8} \end{equation} we have \begin{equation} {\cal P}^{\rm germ}\leftrightarrow P^{\rm germ}\to(\beta, \theta) \leftrightarrow (\alpha,\varphi)\leftrightarrow P^{\rm creat}. \label{3.4.9} \end{equation} Thus \begin{equation} {\cal P}^{\rm germ}\to P^{\rm creat}. \label{3.4.10} \end{equation} \subsection{Regular crossing} Let for $\tau<0$ \begin{equation} \omega_{m}^{\rm creat\:left}=\omega_{m}^{\rm germ\:left} \label{3.5.1} \end{equation} hold. Then it is natural to put for $\tau>0$ \begin{equation} \omega_{m}^{\rm creat\:right}=\omega_{m}^{\rm creat\:left} \equiv\omega_{m}^{\rm creat} \label{3.5.2} \end{equation} and, in view of eqs.(\ref{3.3.9}),(\ref{3.3.10}), \begin{equation} \omega_{m}^{\rm germ\:right}=\omega_{m}^{\rm germ\:left}. \label{3.5.3} \end{equation} Thus, there exists a germ process ${\cal P}_{(-\theta,\theta)} ^{\rm germ}$, such that ${\cal P}_{(-\theta,0)}^{\rm germ\: left}$ and ${\cal P}_{(0,\theta)}^{\rm germ\:right}$ are its restrictions, \begin{equation} \lim_{\tau\to-0}{\cal P}^{\rm germ\:left}(\tau)=\lim_{\tau\to+0} {\cal P}^{\rm germ\:right}(\tau)={\cal P}_{(-\theta,\theta)} ^{\rm germ}(0)= \omega_{m}^{\rm creat}, \label{3.5.4} \end{equation} and there is no jump. The point $\tau=0$ and the process ${\cal P}_{(-\theta,\theta)}^{\rm germ}$ are deterministic. \subsection{Singular crossing and quantum jump} Now let \begin{equation} \omega_{m}^{\rm creat\:left}\ne\omega_{m}^{\rm germ\:left}. \label{3.6.1} \end{equation} There is no possibility for a continuous process ${\cal P}_{(-\theta,\theta)}^{\rm germ}$. Since $\omega_{m0}$ is not determined by the process ${\cal P}_{(-\theta,0)}^{\rm germ\: left}$, we put \begin{equation} \omega_{m0}=\lim_{\tau\to-0}\omega_{m\tau}=\omega_{m}^{\rm germ\:left}. \label{3.6.2} \end{equation} Furthermore, it is natural to put \begin{equation} \omega_{m}^{\rm creat\:right}=\omega_{m0}, \label{3.6.3} \end{equation} so that \begin{equation} \omega_{m}^{\rm creat\:right}=\omega_{m}^{\rm germ\:left}= \lim_{\tau\to-0}\omega_{m\tau}=\omega_{m-0}. \label{3.6.4} \end{equation} We have, by eqs.(\ref{3.4.2}),(\ref{3.4.3}),(\ref{3.4.8}),(\ref{3.6.4}), a quantum jump \begin{equation} P_{-0}^{\rm left}\leftrightarrow \omega_{m-0} \stackrel{\rm jump}{\longrightarrow}\omega_{m+0}^{l} \leftrightarrow P_{+0}^{{\rm right}\:l},\quad l=\pm, \label{3.6.5} \end{equation} with a transition probabilities related to it \begin{equation} w({\cal P}_{(0,\theta)}^{{\rm germ\:right}\:l} /{\cal P}_{(-\theta,0)} ^{\rm germ\:left})={\rm Tr}\{P_{+0}^{{\rm right}\:l} P_{-0}^{\rm left}\},\quad l=\pm. \label{3.6.6} \end{equation} In the case of a regular crossing, eq.(\ref{3.6.6}) gives $w=1$ or 0; this case is an idealized limiting one. Thus, a singular crossing gives rise to a quantum jump. \subsection{Orientability of dynamics and arrow of time} A point which corresponds to a singular crossing is indeterministic. We have for the cardinal numbers related to it \begin{equation} {\rm card^{future\:\alpha}=card^{right\:\alpha}=2>1= card^{left\:\alpha}=card^{past\:\alpha}},\quad \alpha=l=\pm. \label{3.7.1} \end{equation} Thus the dynamics of indeterministic quantum gravity is orientable; it determines the arrow of time given by \begin{equation} t=\tau. \label{3.7.2} \end{equation} We find for the probabilities of subsection 1.8 \begin{equation} w(f/i)={\rm Tr}\{P_{+0}^{{\rm right}\:l} P_{-0}^{\rm left}\},\quad i=1,\quad f=l=\pm, \label{3.7.3} \end{equation} \begin{equation} \sum_{f}w(f/i)={\rm Tr}\{P_{0}P _{-0}^{\rm left}\}= {\rm Tr}\{P_{-0}^{\rm left}\}=1, \label{3.7.4} \end{equation} \begin{equation} \sum_{i}w(f/i)=w(f/1)\leq 1, \label{3.7.5} \end{equation} \begin{equation} w(i/f)=1, \label{3.7.6} \end{equation} so that \begin{equation} w(f/i)\leq w(i/f)=1. \label{3.7.7} \end{equation} \subsection{Nonpredeterminability of the future and reconstructibility of the past} The dynamics developed is indeterministic, therefore the future is not predeterminate and may be forecasted only on a probabilistic level. On the other hand, in view of eqs.(\ref{3.4.10}),(\ref{3.6.4}), we have \begin{equation} {\cal P}_{(0,\theta)}^{\rm germ\:future}\to \omega_{m} ^{\rm create\:right}=\omega_{m}^{\rm germ\:left} \leftrightarrow {\cal P}_{(-\theta,0)}^{\rm germ\:past}, \label{3.8.1} \end{equation} so that \begin{equation} {\cal P}_{(0,\theta)}^{\rm germ\:future}\to {\cal P} _{(-\theta,0)} ^{\rm germ\:past}. \label{3.8.2} \end{equation} Thus, the past is reconstructible uniquely.
proofpile-arXiv_065-405
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
proofpile-arXiv_065-406
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} Mixing phenomena in neutral $B$ meson systems provide us with an important probe of standard model flavordynamics and its interplay with the strong interaction. As is well-known, non-zero off-diagonal elements of the mixing matrix in the flavor basis $\{|B_s\rangle,\, |\bar B_s\rangle\}$ are generated in second order in the weak interaction through `box diagrams'. In the $B_s$ system\footnote{For $B_d$ mesons there is further CKM suppression and their lifetime difference will not be considered here.} the off-diagonal elements obey the pattern \begin{equation} \left|\frac{\Gamma_{12}}{M_{12}}\right|\sim {\cal O}\!\left(\frac{m_b^2} {m_t^2}\right). \end{equation} The mass and lifetime difference between eigenstates are given by (`H' for `heavy', `L' for `light') \begin{equation}\label{delmex} \Delta M_{B_s} \equiv M_H-M_L=2 \,|M_{12}|,\\[0.1cm] \end{equation} \begin{equation}\label{delgex} \Delta \Gamma_{B_s} \equiv \Gamma_L-\Gamma_H = -\frac{2\,\mbox{Re}\,( M_{12}^*\Gamma_{12})}{|M_{12}|}\approx -2 \Gamma_{12}, \end{equation} up to very small corrections (assuming standard model CP violation). Anticipating the magnitudes of the eigenvalues, we have defined both $\Delta M_{B_s}$ and $\Delta\Gamma_{B_s}$ to be positive. Note that the lighter state is CP even and decays more rapidly than the heavier state. The lifetime difference is an interesting quantity in several respects. Contrary to the neutral kaon system, it is calculable by short-distance methods and directly probes the spectator quark dynamics which generates lifetime differences among all $b$ hadrons. If the mass difference $\Delta M_{B_s}$ turns out to be large, the lifetime difference also tends to be large and may well be the first direct observation of mixing for $B_s$ mesons. If $\Delta\Gamma_{B_s}$ is sizeable, CP violation in the $B_s$ system can be observed without flavor-tagging \cite{DUN1}. The following sections summarize the calculation of Ref.~\cite{BBD} and discuss some of the implications of a non-zero $\Delta\Gamma_{B_s}$. \section{Heavy quark expansion of $\Delta\Gamma_{B_s}$} The mass difference is dominated by the top-quark box diagram, which reduces to a local $\Delta B=2$ vertex on a momentum scale smaller than $M_W$. The lifetime difference, on the other hand, is generated by real intermediate states and is not yet local on this scale. But the $b$ quark mass $m_b$ provides an additional short-distance scale that leads to a large energy release (compared to $\Lambda_{QCD}$) into the intermediate states. Thus, at typical hadronic scales the decay is again a local process. The lifetime difference can then be treated by the same operator product expansion that applies to the average $B_s$ lifetime and other $b$ hadrons \cite{BIG}. Summing over all intermediate states, the off-diagonal element $\Gamma_{21}$ of the decay width matrix is given by \begin{equation}\label{g21t} \Gamma_{21}=\frac{1}{2M_{B_s}} \langle\bar B_s|\,\mbox{Im}\,i\!\int\!\!d^4xT\,{\cal H}_{eff}(x) {\cal H}_{eff}(0)|B_s\rangle \end{equation} with \begin{eqnarray}\label{heff} {\cal H}_{eff}&=&\frac{G_F}{\sqrt{2}}V^*_{cb}V_{cs} \big(C_1(\mu) (\bar b_ic_j)_{V-A}(\bar c_js_i)_{V-A} \nonumber\\ && +\,C_2(\mu) (\bar b_ic_i)_{V-A}(\bar c_js_j)_{V-A}\big). \end{eqnarray} Cabibbo suppressed and penguin operators in $ {\cal H}_{eff}$ have not been written explicitly. In leading logarithmic approximation, the Wilson coefficients are given by $C_{2,1}=(C_+\pm C_-)/2$, where \begin{equation} C_+(\mu)\!=\!\left[\frac{\alpha_s(M_W)}{\alpha_s(\mu)}\right]^{6/23}\!\!\!\! C_-(\mu)\!=\!\left[\frac{\alpha_s(M_W)}{\alpha_s(\mu)}\right]^{-12/23} \end{equation} \noindent and $\mu$ is of order $m_b$. The heavy quark expansion expresses $\Delta\Gamma_{B_s}$ as a series in local $\Delta B=2$-operators. In the following we keep $1/m_b$-corrections to the leading term in the expansion. Keeping these terms fixes various ambiguities of the leading order calculation, such as whether the quark mass $m_b$ or meson mass $M_{B_s}$ should be used, and establishes the reliability of the leading order expression obtained in Ref.~\cite{LO,VOL}. Compared to the `exclusive approach' pursued in Ref.~\cite{ALE93} that adds the contributions to $\Delta\Gamma_{B_s}$ from individual intermediate states, the inclusive approach is model-independent. The operator product expansion provides a systematic approximation in $\Lambda_{QCD}/m_b$, but it relies on the assumption of `local duality'. The accuracy to which one should expect duality to hold is difficult to quantify, except for models \cite{CHI96} and eventually by comparison with data. We shall assume that duality violations will be less than 10\% for $\Delta\Gamma_{B_s}$. To leading order in the heavy quark expansion, the long distance contributions to $\Delta\Gamma_{B_s}$ are parameterized by the matrix elements of two dimension six operators \begin{eqnarray}\label{qqs} Q &=& (\bar b_is_i)_{V-A}(\bar b_js_j)_{V-A},\\[0.1cm] Q_S\!\!&=& (\bar b_is_i)_{S-P}(\bar b_js_j)_{S-P} \end{eqnarray} between a $\bar{B}_s$ and $B_s$ state. We write these matrix elements as \begin{eqnarray}\label{meqs} \langle Q\rangle &=& \frac{8}{3}\, f^2_{B_s}M^2_{B_s}\,B, \\[0.1cm] \langle Q_S\rangle\!\! &=& - \frac{5}{6} \,f^2_{B_s}M^2_{B_s} \frac{M^2_{B_s}}{(m_b+m_s)^2}\,B_S, \end{eqnarray} where $f_{B_s}$ is the $B_s$ decay constant. The `bag' parameters $B$ and $B_S$ are defined such that $B=B_S=1$ corresponds to factorization. $B$ also appears in the mass difference, while $B_S$ is specific to $\Delta\Gamma_{B_s}$. The matrix elements of these operators are not independent of $m_b$. Their $m_b$-dependence could be extracted with the help of heavy quark effective theory. There seems to be no gain in doing so, since the number of independent nonperturbative parameters is not reduced even at leading order in $1/m_b$ and since we work to subleading order in $1/m_b$ even more parameters would appear. The matrix elements of the local $\Delta B=2$-operators should therefore be computed in `full' QCD, for instance on the lattice. Including $1/m_b$-corrections, the width difference is found to be \begin{eqnarray}\label{tres} \Delta\Gamma_{B_s} &=& \frac{G^2_F m^2_b}{12\pi M_{B_s}}(V^*_{cb}V_{cs})^2 \sqrt{1-4z} \nonumber\\ &&\hspace*{-1.7cm}\,\cdot\bigg[\left((1-z)K_1+ \frac{1}{2}(1-4z)K_2\right)\langle Q \rangle\\ &&\hspace*{-1.7cm}\,+\,(1+2z)\left(K_1-K_2\right)\langle Q_S\rangle + \hat{\delta}_{1/m} + \hat{\delta}_{rem} \bigg], \nonumber \end{eqnarray} where $z=m^2_c/m^2_b$ and \begin{equation}\label{k1k2} K_1=N_c C^2_1+2C_1 C_2\qquad K_2=C^2_2 . \end{equation} The $1/m_b$-corrections are summarized in \begin{eqnarray}\label{oneoverm} \hat{\delta}_{1/m} &=& (1+2 z)\Big[K_1\,(-2\langle R_1\rangle -2\langle R_2\rangle) +\,K_2\,(\langle R_0\rangle -2\langle \tilde{R}_1 \rangle -2\langle \tilde{R}_2\rangle)\Big] \nonumber\\ &&-\,\frac{12 z^2}{1-4 z}\Big[K_1\,(\langle R_2\rangle +2\langle R_3\rangle) +\,K_2\,(\langle \tilde{R}_2\rangle +2\langle \tilde{R}_3\rangle)\Big] . \end{eqnarray} The operators $R_i$ and $\tilde{R_i}$ involve derivatives on quark fields or are proportional to the strange quark mass $m_s$, which we count as $\Lambda_{QCD}$. For instance, \begin{eqnarray} \label{r0qt} R_1\!&=&\!\frac{m_s}{m_b}(\bar b_is_i)_{S-P}(\bar b_js_j)_{S+P},\\ \label{rrt2} R_2\!&=&\!\frac{1}{m^2_b}(\bar b_i {\overleftarrow D}_{\!\rho} D^\rho s_i)_{V-A}( \bar b_j s_j)_{V-A}. \label{rrt3} \end{eqnarray} The complete set can be found in Ref.~\cite{BBD}. Operators with gluon fields contribute only at order $(\Lambda_{QCD}/m_b)^2$. Since the matrix elements of the $R_i$, $\tilde{R}_i$ are $1/m_b$-suppressed compared to those of $Q$ and $Q_S$, we estimate them in the factorization approximation, assuming factorization at a scale of order $m_b$ (A smaller scale would be preferable, but would require us to calculate the anomalous dimension matrix.). Then all matrix elements can be expressed in terms of quark masses and the $B_s$ mass and decay constant. No new nonperturbative parameters enter at order $1/m_b$ in this approximation. The term $\hat{\delta}_{rem}$ denotes the contributions from Cabibbo-suppressed decay modes and pengiun operators. They can be estimated \cite{BBD} to be below $\pm 3\%$ and about $-5\%$, respectively, relative to the leading order contribution. We neglect this term in the following numerical analysis. \section{Numerical estimate} \begin{table}[t] \addtolength{\arraycolsep}{-0.01cm} \renewcommand{\arraystretch}{1.3} \caption{\label{table1} Dependence of $a$, $b$ and $c$ on the $b$-quark mass (in GeV) and renormalization scale for fixed values of all other short-distance parameters. The last column gives $(\Delta\Gamma/\Gamma)_{B_s}$ for $B=B_S=1$ (at given $\mu$), $f_{B_s}=210\,$MeV.} $$ \begin{array}{|c|c||c|c|c|c|} \hline m_b & \mu & a & b & c & (\Delta\Gamma/\Gamma)_{B_s} \\ \hline\hline 4.8 & m_b & 0.009 & 0.211 & -0.065 & 0.155 \\ \hline 4.6 & m_b & 0.015 & 0.239 & -0.096 & 0.158 \\ \hline 5.0 & m_b & 0.004 & 0.187 & -0.039 & 0.151 \\ \hline 4.8 & 2 m_b & 0.017 & 0.181 & -0.058 & 0.140 \\ \hline 4.8 & m_b/2 & 0.006 & 0.251 & -0.076 & 0.181 \\ \hline \end{array} $$ \end{table} It is useful to separate the dependence on the long-distance parameters $f_{B_s}$, $B$ and $B_S$ and write $(\Delta\Gamma/\Gamma)_{B_s}$ as \begin{equation} \left(\frac{\Delta\Gamma}{\Gamma}\right)_{B_s} = \Big[a B + b B_S + c\Big]\left(\frac{f_{B_s}}{210\,\mbox{MeV}}\right)^2, \end{equation} where $c$ incorporates the explicit $1/m_b$-corrections. In the numerical analysis, we express $\Gamma_{B_s}$ as the theoretical value of the semileptonic width divided by the semileptonic branching ratio. The following parameters are kept fixed: $m_b-m_c=3.4\,$GeV, $m_s=200\,$MeV, $\Lambda^{(5)}_{LO}=200\, $MeV, $M_{B_s}=5.37\,$GeV, $B(B_s\to X e\nu)=10.4\%$. Then $a$, $b$ and $c$ depend only on $m_b$ and the renormalization scale $\mu$. For some values of $m_b$ and $\mu$, the coefficients $a$, $b$, $c$ are listed in Tab.~\ref{table1}. For a central choice of parameters, which we take as $m_b=4.8\,$GeV, $\mu=m_b$, $B=B_S=1$ and $f_{B_s}=210\,$MeV, we obtain $(\Delta\Gamma/\Gamma)_{B_s} = 0.220 - 0.065 = 0.155$, where the leading term and the $1/m_b$-correction are separately quoted. We note that the $V-A$ `bag' parameter $B$ has a very small coefficient and is practically negligible. The $1/m_b$-corrections are not small and decrease the prediction for $\Delta\Gamma_{B_s}$ by about 30\%. The largest theoretical uncertainties arise from the decay constant $f_{B_s}$ and the second `bag' parameter $B_S$. In the large-$N_c$ limit, one has $B_S=6/5$, while estimating $B_S$ by keeping the logarithmic dependence on $m_b$ (but not $1/m_b$-corrections as required here for consistency) and assuming factorization at the scale $1\,$GeV gives \cite{VOL} $B_S=0.88$. $B_S$ has never been studied by either QCD sum rules or lattice methods. In order to estimate the range of allowed $\Delta\Gamma_{B_s}$ conservatively, we vary $B_S=1\pm0.3$, $f_{B_s}=(210\pm30)\,$MeV and obtain \begin{equation}\label{dgnum1} \left(\frac{\Delta\Gamma}{\Gamma}\right)_{B_s} = 0.16^{+0.11}_{-0.09}. \end{equation} This estimate could be drastically improved with improved knowledge of $B_S$ and $f_{B_s}$. \section{Measuring $\Delta\Gamma_{B_s}$} In principle, both $\Gamma_L$ and $\Gamma_H$ can be measured by following the time-dependence of flavor-specific modes \cite{DUN1}, such as $\bar{B}_s\to D_s l\nu$, given by \begin{equation} e^{-\Gamma_H t}+e^{-\Gamma_L t} . \end{equation} In practice, this is a tough measurement. Alternatively, since the average $B_s$ lifetime is predicted~\cite{BBD} to be equal to the $B_d$ lifetime within $1\%$, it is sufficient to measure either $\Gamma_L$ or $\Gamma_H$. The two-body decay $B_s\to D_s^+ D_s^-$ has a pure CP even final state and measures $\Gamma_L$. Since $D^0$ and $D^{\pm}$ do not decay into $\phi$ as often as $D_s$, the $\phi\phi X$ final state tags a $B_s$-enriched $B$ meson sample, whose decay distribution informs us about $\Gamma_L$. A cleaner channel is $B_s\to J/\psi\phi$, which has both CP even and CP odd contributions. These could be disentangled by studying the angular correlations \cite{DIG96}. In practice, this might not be necessary, as the CP even contribution is expected \cite{ALE93} to be dominant by more than an order of magnitude. In any case, the inequality \begin{equation} \Gamma_L \geq 1/\tau(B_s\to J/\psi\phi) \end{equation} holds. CDF \cite{MES} has fully reconstructed 58 $B_s\to J/\psi\phi$ decays from run Ia+Ib and determined $\tau(B_s\to J/\psi\phi)= 1.34^{+0.23}_{-0.19}\pm 0.05\,$ps. Together with $\tau(B_d)=1.54\pm 0.04\,$ps, assuming equal average $B_d$ and $B_s$ lifetimes, this yields \begin{equation} \left(\frac{\Delta\Gamma}{\Gamma}\right)_{B_s}\geq 0.3\pm 0.4, \end{equation} which still fails to be significant. In the Tevatron run II, as well as at HERA-B, one expects $10^3-10^4$ reconstructed $J/\psi\phi$, which will give a precise measurement of $\Delta\Gamma_{B_s}$. \section{Implications of non-zero $\Delta\Gamma_{B_s}$} \subsection{CKM elements} Once $\Delta\Gamma_{B_s}$ is measured (possibly before $\Delta M_{B_s}$ is measured!), an alternative route to obtain the mass difference could use this measurement combined with the theoretical prediction for $(\Delta M/ \Delta\Gamma)_{B_s}$ \cite{DUN1,BP}. The decay constant $f_{B_s}$ drops out in this ratio, as well as the dependence on CKM elements, since $|(V_{cb}V_{cs})/(V_{ts}V_{tb})|^2= 1\pm 0.03$ by CKM unitarity. However, the dependence on long-distance matrix elements does not cancel even at leading order in $1/m_b$ and the prediction depends on the ratio of `bag' parameters $B_S/B$, which is not very well-known presently. We obtain $\Delta\Gamma/\Delta M=(5.6\pm 2.6)\cdot 10^{-3}$, where the largest error ($\pm 2.3$) arises from varying $B_S/B$ between 0.7 and 1.3. When lattice measurements yield an accurate value of $B_S/B$ as well as control over the $SU(3)$ flavor-symmetry breaking in $B f_B^2$, the above indirect determination of $\Delta M_{B_s}$ in conjunction with the measured mass difference in the $B_d$ system provides an alternative way of determining the CKM ratio $|V_{ts}/V_{td}|$, especially if the latter is around its largest currently allowed value. In contrast, the ratio $\Gamma(B\to K^*\gamma)/\Gamma(B\to\{\varrho,\omega\}\gamma)$ is best suited for extracting small $|V_{ts}/V_{td}|$ ratios, provided the long distance effects can be sufficiently well understood. \subsection{CP violation} The existence of a non-zero $\Delta\Gamma_{B_s}$ allows the observation of mixing-induced CP asymmetries without tagging the initial $B_s$ or $\bar{B}_s$ \cite{DUN1,DF}. These measurements are difficult, but the gain in statistics, when tagging is obviated, makes them worthwhile to be considered. The mass difference drops out in the time dependence of untagged samples, which is given by \begin{equation} A_+ (e^{-\Gamma_L t}+e^{-\Gamma_H t}) + A_-(e^{-\Gamma_L t}-e^{-\Gamma_H t}). \end{equation} $A_-$ carries CKM phase information even in the absence of direct CP violation. In combination with an analysis of angular distributions, a measurement of the CKM angle $\gamma$ from exclusive $B_s$ decays governed by the $\bar{b}\to c\bar{c}\bar{s}$ or $\bar{b}\to\bar{c}u\bar{s}$ transition can be considered \cite{DF}. \section*{Acknowledgements} I am grateful to my collaborators G.~Buchalla and I.~Dunietz for sharing their insights into problems related to this work with me. \section*{References}
proofpile-arXiv_065-407
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} Since the early days of quantum field theory (QFT) 1+1 dimensional models have attracted much attention. They have been extremely valuable to develop general ideas and intuition about the structure of QFT. The eldest and perhaps most popular of these 1+1 D models bear the names of Thirring \cite{Thirring} -- Dirac fermions interacting with a local current-current interaction -- and Schwinger \cite{Schwinger} -- quantum electrodynamics with fermions. The models originated in particle physics and therefore, in order to have Lorentz invariance, were considered mainly on infinite space ${\bf R}$ (see e.g.\ \cite{CHu,CW,CRW} and references therein). Then one has to deal with infrared (infinite space volume) divergences in addition to singularities coming from the ultraviolet (short distances). In case the fermions are massless, both models are soluble \cite{Schwinger,Klaiber,swieca} and a very detailed picture of their properties can be obtained. Another related model originated from solid state physics and is due to Luttinger \cite{Luttinger} -- massless Dirac fermions on spacetime $S^1\times{\bf R}$ interacting with a non-local current-current interaction (Lorentz invariance is nothing natural to ask for in solid state physics). The Luttinger model shows that an interacting fermion system in one space dimension need not behave qualitatively similar to free fermions but rather has properties similar to a boson system. Such behaviour is generic for 1+1 D interacting fermion models and is denoted as Luttinger liquid in solid state physics, in contrast to Landau liquids common in 3+1 D. To consider the Luttinger model on compact space has the enormous technical advantage that infra red (IR) problems are absent, and one can concentrate on the short distance (UV) properties which are rather simple due to the non-locality of the interaction. In fact, this allows a construction of the interacting model on the Fock space of {\em free} fermions \cite{MattisLieb,HSU,CH} and one directly can make use of mathematical results from the representation theory of the affine Kac-Moody algebras. Such an approach was recently given for QCD with massless fermions \cite{LS3}. As shown by Manton \cite{Manton}, the Schwinger model on compact space $S^1$ allows a complete understanding of the UV divergences and anomalies and their intriguing interplay with gauge invariance and vacuum structure. In the present paper we study the extension of the Luttinger model obtained by coupling it to a dynamical electromagnetic field. For vanishing Luttinger (4-point) interaction our model therefore reduces to the Schwinger model as studied by Manton \cite{Manton}, and for vanishing electric charge to the Luttinger model \cite{MattisLieb}. Since our approach is in Minkowski space and provides a direct construction of the field-- and observable algebras of the model on a physical Hilbert space, it is conceptually quite different from the path integral approach, and we believe it adds to the physical understanding of these models. The plan of the paper is as follows. In Section 2 the construction of the model is given. To fix notation, we first summarize the classical Hamiltonian formalism. We then construct the physical Hilbert space and discuss the non-trivial implications of anomalies (Schwinger terms) and gauge invariance. In Section 3 the model is solved by bosonization, and a method for calculating all Green functions is explained. As an example the equal time 2-point functions are given. In Section 4 we comment on regularization and renormalization in our setting. We discuss the limit to the Thirring-Schwinger model where the 4-point interaction becomes local and space infinite. We end with a short summary in Section 5. A summary of the mathematical results needed and some details of calculations are deferred to the appendix. \section{Constructing the model} \subsection{Notation} Spacetime is the cylinder with $x=x^1\in \Lambda\equiv [-L/2,L/2]$ the spatial coordinate and $t=x^0\in{\bf R}$ time. We have one Dirac Fermion field $\psi_{\sigma}(\vec{x})$ and one Photon field $A_\nu(\vec{x})$ (here and in the following, $\sigma,\sigma'\in\{+,-\}$ are spin indices, $\mu,\nu\in\{0,1\}$ are spacetime indices, and $\vec{x}=(t,x),\vec{y}=(t',y)$ are spacetime arguments). The action defining the Luttinger-Schwinger model is\footnote{unless otherwise stated, repeated indices are summed over throughout the paper}$^,$ \footnote{$\partial_\nu\equiv\partial/\partial x^\nu$; our metric tensor is $g_{\mu\nu}=diag(1,-1)$} \begin{eqnarray} \label{1} {\cal S} = \int d^2\vec{x}\left(-\frac{1}{4}F_{\mu\nu}(\vec x)F^{\mu\nu}(\vec x) + \bar\psi(\vec x) \gamma^\nu\left(-{\rm i}\partial_\nu + e A_\nu(\vec x)\right)\psi(\vec x)\right) \nonumber \\ \nopagebreak - \int d^2\vec{x}\int d^2\vec{y}\, j_\mu (\vec x)v(\vec x -\vec y) j^\mu(\vec y) \end{eqnarray} where $F_{\mu\nu}=\partial_\mu A_\nu-\partial_\nu A_\mu$ and $\gamma^\nu\equiv (\gamma^\nu)_{\sigma\sigma'}$ are Dirac matrices which we take as $\gamma^0=\sigma_1$ and $\gamma^1={\rm i}\sigma_2$, and $\gamma_5=-\gamma^0\gamma^1=\sigma_3$ ($\sigma_i$ are Pauli spin matrices). As usual, the fermion currents are $ j_\nu = \bar\psi\gamma_\nu\psi, $ and we assume the 4--point interaction to be instantaneous (local in time) \begin{equation} v(\vec x-\vec y) = \delta(t-t')V(x-y) \end{equation} where the interaction potential is parity invariant, $V(x)=V(-x)$. As in case of the Luttinger model \cite{HSU} we will also have to assume that this potential is not `too strong', or more precisely that the Fourier coefficients \alpheqn \begin{equation} \label{condition} W_k = \frac{1}{8\pi}\int_{\Lam}{\rm d}{x}\, V(x)\ee{-{\rm i} kx} = W_{-k}= W_k^*,\quad k\frac{L}{2\pi} \in{\bf Z} \end{equation} of the potential obey the conditions \begin{equation} \label{condition1} -1-\frac{e^2}{\pi k^2} < W_k < 1\quad\forall k\quad \mbox{ and }\quad \sum_k |kW_k^2|<\infty . \end{equation} \reseteqn {}From the action \Ref{1} we obtain the canonical momenta $\Pi_{A_0(x)} \simeq 0$, $\Pi_{A_1(x)} = F_{01}(x) = E(x)$ etc.\ (here and in the following, we set $t=0$ and make explicit the dependence on the spatial coordinate only) resulting in the Hamiltonian ($\psi^*\equiv \bar\psi\gamma^0$) \begin{eqnarray} \label{4} H=\int_{\Lam} {\rm d}{x} \left( \frac{1}{2} E(x)^2 + \psi^*(x)\gamma_5\left(-{\rm i}\partial_1 + e A_1(x)\right) \psi(x)\right) + 4\int_{\Lam} {\rm d}{x} {\rm d}{y} \, \rho^+(x)V(x-y)\rho^-(y), \end{eqnarray} and the Gauss' law \begin{equation} \label{5} G(x) = -\partial_1 E(x) + e \rho(x) \simeq 0 \, . \end{equation} We introduced chiral fermion currents \begin{equation} \rho^\pm(x) = \psi^*(x)\frac{1}{2}(1\pm\gamma_5)\psi(x) \label{chiralcu} \end{equation} so that fermion charge-- and momentum density $\rho=j^0$ and $j=j^1$ can be written as \begin{eqnarray} \rho(x) &=& \rho^+(x) + \rho^-(x)\nonumber \\ \nopagebreak j(x) &=& \rho^+(x) - \rho^-(x). \end{eqnarray} \subsection{Observables} \label{obs} The observables of the model are all gauge invariant operators. They leave invariant physical states. The ground state expectation values of these operators are the Green functions we are interested in. For later reference we write down the action of static gauge transformations i.e.\ differentiable maps $\Lambda\to{\rm U}(1), x\mapsto \ee{{\rm i} \alpha(x)}$, \begin{eqnarray} \label{gaugetrafo} \psi_\sigma(x)&\to& \ee{{\rm i} \alpha(x)}\psi_\sigma(x) \nonumber \\ \nopagebreak A_1(x) &\to& A_1(x) - \frac{1}{e}\frac{\partial \alpha(x)}{\partial x} \\ E(x) &\to& E(x)\, . \nonumber \end{eqnarray} These obviously leave our Hamiltonian and Gauss' law invariant. We note that every gauge transformation can be decomposed into a {\em small} and a {\em large} gauge transformation, $ \alpha(x)=\alpha_{small}(x) + \alpha_{large}(x), $ where \begin{equation} \alpha_{large}(x) = n\frac{2\pi x}{L} \quad (n\in{\bf Z}), \quad \alpha_{small}\left(-\frac{L}{2}\right)= \alpha_{small}\left(\frac{L}{2}\right) \end{equation} with $n=\frac{\alpha(L/2)-\alpha(-L/2)}{2\pi L}$. The large gauge transformations correspond to $\Pi_1(S^1)={\bf Z}$ and play an important role in the following, as expected from general arguments \cite{Jackiw}. It is important to note that Gauss' law \Ref{5} requires physical states only to be invariant under small (but {\em not} under large) gauge transformations. All gauge invariant objects which one can construct from $A_1(x)$ (at fixed time) are functions of \begin{equation} Y= \frac{1}{2\pi} \int_{\Lambda}{\rm d} y A_1(y) \, . \end{equation} In fact, $Y$ above is only invariant with respect to small gauge transformations and changes by multiples of $1/ e$ under the large ones. Thus the quantity which is invariant under all gauge transformations is $\ee{{\rm i} 2\pi e Y}$ which is equal to the Wilson line (holonomy) \begin{equation} \label{Wilson} W[A_1] = \ee{{\rm i} e\int_{\Lambda}{\rm d} y A_1(y)} \: . \end{equation} The fermion fields are not gauge invariant, but by attaching parallel transporters to them one obtains field operators \begin{equation} \label{chi} \chi_\sigma(x) = \ee{{\rm i} e\int_r^x{\rm d} y A_1(y) }\psi_\sigma(x)\, , \quad r\in\Lambda \end{equation} which obviously are invariant under all (small and large) gauge transformations \Ref{gaugetrafo} with $\alpha(r)=0$; $r$ is a spatial point which we can choose arbitrarily. Note that these fields also obey CAR but are {\em not} antiperiodic: they obey $\chi_\sigma(L/2)= -W[A_1]\chi_\sigma(-L/2)$ where $W[A_1]$ is the Wilson line above. Bilinears of these operators are the meson operators $$ M_{\sigma\sigma'}(x,y) = \chi^*_\sigma(x)\chi_{\sigma'}(y) \, . $$ These are invariant under all static gauge transformations and thus can be used as building blocks of the Green functions we are interested in. \subsection{The quantum model} In the following we find it convenient to work in Fourier space. We introduce the following useful notation. Fourier space for even (periodic) functions is \alpheqn \begin{equation} \label{a} \Lambda^*\equiv \left\{\left. k=\frac{2\pi}{L} n \right| n\in{\bf Z}\right\}\quad \, . \end{equation} As we use fermions with odd (anti--periodic) boundary conditions we also need \begin{equation} \Lambda^*_{odd}\equiv \left\{\left. k=\frac{2\pi}{L} \left(n+\frac{1}{2}\right) \right| n\in{\bf Z}\right\}. \end{equation} {}For functions $\hat f$ on Fourier space we write \begin{equation} \hat{\int}_{\dLam} \hat{\rm d} k \hat f(k) \equiv \sum_{k\in\Lambda^*} \frac{2\pi}{L} \hat f(k) \end{equation} \reseteqn and similarly for $\Lambda^*_{odd}$ (we will use the same symbols $\delta$ and $\hat\delta$ also in the latter case). Then the appropriate $\delta$-function satisfying $\hat{\int}_{\dLam}\hat{\rm d} q\, \hat{\delta}(k-q)\hat f(q)=\hat f(k)$ is $\hat{\delta}(k-q) \equiv \frac{L}{2\pi} \delta_{k,q}$. {}For the Fourier transformed operators we use the following conventions, \alpheqn \begin{equation} \label{10a} \hat\psi^{}_{\sigma}(q) = \int_{\Lam} \frac{{\rm d}{x}}{\sqrt{2\pi}} \psi^{}_{\sigma}(x) \ee{-{\rm i} qx},\quad \hat\psi^{*}_{\sigma}(q)=\hat\psi^{}_{\sigma}(q)^* \quad (q\in\Lambda^*_{odd}) \end{equation} (as mentioned, we use anti--periodic boundary conditions for the fermions), \begin{equation} \label{A1} \hat A_1(k) = \int_{\Lam} \frac{{\rm d}{x}}{2\pi} A_1(x) \ee{-{\rm i} kx} \quad (k\in\Lambda^*) \end{equation} and in the other cases \begin{equation} \label{other} \hat Y(k) = \int_{\Lam} {\rm d}{x}\, Y(x) \ee{-{\rm i} kx}\quad (k\in\Lambda^*)\quad \mbox{ for $Y=E,\rho^{\pm},\rho,j,V$} \end{equation} \reseteqn Following \cite{HSU} we also find it convenient to introduce $W_k=\hat V(k)/8\pi$ (cf.\ \Ref{condition}). With that the non-trivial C(A)CR in Fourier space are \begin{eqnarray} \label{fcacr} \ccr{\hat A_1(p)}{\hat E(k)} &=& {\rm i}\hat{\delta}(k+p) \nonumber \\ \nopagebreak \car{\hat\psi_{\sigma}(q)}{\hat\psi^*_{\sigma'}(q')} &=& \delta_{\sigma\sigma'}\hat{\delta}(q-q') . \end{eqnarray} The essential physical requirement determining the construction of the model and implying a non-trivial quantum structure is positivity of the Hamiltonian on the physical Hilbert space. It is well-known that it forces one to use a non-trivial representation of the field operators of the model. The essential simplification in (1+1) (and not possible in higher) dimensions is that one can use a quasi-free representation for the fermion field operators corresponding to ``filling up the Dirac sea'' associated with the {\em free} fermion Hamiltonian, and for the photon operators one can use a naive boson representation. This will be verified for our model for the class of potentials $V$ obeying (\ref{condition},b). So the full Hilbert space of the model is ${\cal H} = {\cal H}_{\rm Photon}\otimes {\cal H}_{\rm Fermion}$. For ${\cal H}_{\rm Photon}$ we take the boson Fock space generated by boson field operators $b^{*}(k)$ obeying CCR \alpheqn \begin{equation} \ccr{b(k)}{b^*(p)} = \hat{\delta}(k-q)\quad\mbox{etc.} \end{equation} and a vacuum\footnote{Note that the term ``vacuum'' here and in the following does {\em not} mean that this state has anything to do with the ground state of the model; it is just one convenient state from which all other states in the Hilbert space can be generated by applying the field operators.} $\Omega_{\rm P}$ such that \begin{equation} b(k)\Omega_{\rm P} = 0 \quad \forall k\in\Lambda^* . \end{equation} \reseteqn We then set \alpheqn \begin{equation} \label{photon} \hat A_1(k) = \frac{1}{s}\left(b(k) + b^*(k)\right) \quad \hat E(k) = -\frac{{\rm i} s}{2}\left(b(k)-b^*(k)\right) \end{equation}\reseteqn where $ s^4 = \pi e^2 $ (the reason for choosing this factor $s$ will become clear later). We will use below normal ordering $\xx\cdots\xx$ of bilinears in the Photon field operators with respect to the vacuum $\Omega_{\rm P}$, for example $\xx b(k)b^*(p)\xx\, = b^*(p)b(k)$. {}For ${\cal H}_{\rm Fermion}$ we take the Fermion Fock space with vacuum $\Omega_{\rm F}$ such that \begin{eqnarray} \label{11} \frac{1}{2}(1 \pm \gamma_5) \hat\psi(\pm q) \,\Omega_{\rm F} &=& \frac{1}{2}(1 \mp \gamma_5) \hat\psi^*(\mp q) \,\Omega_{\rm F} =0 \quad \forall q > 0\, . \end{eqnarray} The presence of the Dirac sea requires normal-ordering $\normal{\cdots}$ of the Fermion bilinears such as $\hat H_0=\hat{\int}_{\dLamF}\hat{\rm d} q \normal{q\,\hat\psi^*(q)\gamma_5\hat \psi(q)}$ and $\hat \rho_\pm$ (\ref{chiralcu}). This modifies their naive commutator relations following from the CAR as Schwinger terms show up \cite{GLrev,CR,A}. In our case, the relevant commutators are: \begin{eqnarray} \ccr{\hat\rho^\pm(k)}{\hat\rho^{\pm}(p)}&=& \pm k\hat{\delta}(k+p)\,, \nonumber\\ \ccr{\hat\rho^\pm(k)}{\hat\rho^{\mp}(p)} &=& 0 \label{12} \, , \\ \ccr{\hat H_0}{\hat\rho^\pm(k)} &=& \pm k \hat\rho^\pm(k) \nonumber \, . \end{eqnarray} We note that \begin{equation} \label{vacF} \hat\rho^+(k)\Omega_{{\rm F}} = \hat\rho^-(-k)\Omega_{{\rm F}} = 0\quad\forall k>0 \end{equation} which together with \Ref{12} shows that the $\hat\rho^+(k)$ (resp. $\hat\rho^-(k)$) give a highest (resp. lowest) weight representation of the Heisenberg algebra. We can now write the Gauss' law operators in Fourier space as \begin{equation} \label{14} \hat G(k) = -{\rm i} k\hat E(k) + e\hat\rho(k), \end{equation} so eqs. \Ref{12} imply \[ \ccr{\hat G(k)}{\hat\rho^\pm(p)} = \pm k e\hat{\delta}(k+p). \] Due to the presence of the Schwinger terms, these Fermion currents no longer commute with the Gauss' law generators, hence they are not gauge invariant and no observables of the model. To obtain Fermion currents obeying the classical relations (without Schwinger terms), we note that $ \ccr{\hat G(k)}{\hat A_1(p)} = k\hat{\delta}(k+p),$ hence the operators \begin{equation} \label{16} \tilde\rho^\pm(k) \equiv \hat\rho^\pm(k) \pm e\hat A_1(k) \end{equation} commute with the Gauss law generators and are thus the observables of the model corresponding to the chiral Fermion currents on the quantum level. Recalling the normalization is only unique up to finite terms, it is natural to regard the $\tilde \rho^\pm(k)$ as the fermion currents obtained by a {\em gauge covariant normal ordering} preserving the classical transformation properties under gauge transformations. Indeed, these currents can be shown to be identical to those obtained by the gauge invariant point splitting method. Similarly, the naive Hamiltonian $\hat H=\hat H_1+\hat H_2$, \begin{eqnarray*} \hat H_1 &=& \hat H_0 +\hat{\int}_{\dLam}\hat{\rm d} k \xx \left( \frac{1}{4\pi}\hat E(k)\hat E(-k) + e\hat A_1(k)\hat j(-k)\right)\xx \\ \hat H_2 &=& \hat{\int}_{\dLam}\hat{\rm d} k \, \hat\rho^+(k)W_k \hat\rho^-(-k) \end{eqnarray*} is not gauge invariant: $\hat H_1$ -- which is the naive Hamiltonian of the Schwinger model -- obeys \[ \ccr{\hat G(k)}{\hat H_1} = 2ke^2 \hat A_1(k) \] and therefore becomes gauge invariant only after adding a photon mass term \cite{Manton} \[ \hat{\int}_{\dLam}\hat{\rm d} k \, e^2 \hat A_1(k)\hat A_1(-k) \] (note that in position space this mass term has the usual form $\frac{e^2}{2\pi}\int_{\Lam}{\rm d}{x}A_1(x)^2$, i.e.\ the photon mass--squared is $e^2/\pi$). Also the Luttinger--interaction term $\hat H_2$ becomes gauge invariant only if one replaces the non--gauge invariant currents $\hat\rho^\pm$ by the gauge invariant $\tilde\rho^\pm$ ones. Thus we obtain the gauge invariant Hamiltonian of the Luttinger--Schwinger model as follows, \begin{eqnarray} H= \hat H_0 + \hat{\int}_{\dLam}\hat{\rm d} k \xx\left( \frac{1}{4\pi} \hat E(k) \hat E(-k) + e\hat A_1(k) \hat j(-k) + e^2 \hat A_1(k)\hat A_1(-k) \right.\nonumber \\ \nopagebreak\left.\frac{}{}\!\!\!+ \left[\hat\rho^+(k)+e\hat A_1(k)\right] W_k\left[\hat\rho^-(-k)-e\hat A_1(-k)\right]\right)\xx. \end{eqnarray} We can now explain the choice (\ref{photon},b) for the representation of the Photon field: the factor $s$ is determined such that the free Photon Hamiltonian is equal to $\hat{\int}_{\dLam}\hat{\rm d} k \sqrt{\frac{e^2}{\pi}}\, b^*(k)b(k)$. \subsection{Bosonization} Kronig's identity\footnote{in the modern literature this is often referred to as (special case of the) Sugawara construction} allows us to rewrite the free Hamiltonian as $ \hat H_0 = \frac{1}{2}\hat{\int}_{\dLam}\hat{\rm d} k \, \xx \left( \hat\rho^+(k)\hat\rho^+(-k) \right.$ $\left.+ \hat\rho^-(k) \hat\rho^-(-k) \right) \xx $ (cf.\ Appendix A for the precise definition of normal ordering; for simplicity of notation we do not distinguish the normal ordering symbol for the photon fields and the fermion currents). With that, it follows from eq.\ \Ref{16} that \begin{equation} \label{20} H = \hat{\int}_{\dLam}\hat{\rm d} k\, \xx \left(\frac{1}{2}\left( \tilde\rho^+(k)\tilde\rho^+(-k) + \tilde\rho^-(k) \tilde\rho^-(-k) \right) + \frac{1}{4\pi}\hat E(k) \hat E(-k) + \tilde\rho^+(k) W_k \tilde\rho^-(-k)\right)\xx \end{equation} which is now explicitly gauge invariant. \section{Solution of the model} \subsection{Gauge Fixing} The only gauge invariant degree of freedom of the Photon field at fixed time is the holonomy $\int_\Lambda{\rm d}{x} A_1(x)$ and one can gauge away all Fourier modes $\hat A_1(k)$ of the gauge field except the one for $k=0$. Thus we can impose the gauge condition \begin{equation} \label{22} \hat A_1(k) = \delta_{k,0} Y \makebox{, \qquad } A_1(x)=\frac{2\pi}{L}Y \end{equation} and solve the Gauss' law $\hat G(k)\simeq 0$ (cf.\ eq.\ \Ref{14}) as \begin{equation} \hat E(k) \simeq \frac{e\hat\rho(k)}{{\rm i} k} \quad {\mbox{for $k\neq 0$}}. \end{equation} This determines all components of $E$ except those conjugate to $Y$: $\hat E(0) = \frac{L}{2\pi} \frac{\partial}{{\rm i} \partial Y}$. After that we are left with the ($k=0$)--component of Gauss' law, {\em viz.} \begin{equation} eQ_0\simeq 0,\quad Q_0 = \hat\rho(0) = \hat\rho^+ (0) + \hat\rho^- (0)\, . \end{equation} Inserting this into \Ref{20}, gives the Hamiltonian of the model on the physical Hilbert space ${\cal H}_{\rm phys}={\cal L}^2({\bf R},{\rm d}Y)\otimes {\cal H}'_{\rm Fermion}$ (where ${\cal H}'_{\rm Fermion}$ is the zero charge sector of the fermionic Fock space): \begin{eqnarray} \label{24} H = -\frac{L}{8\pi^2} \frac{\partial^2}{\partial Y^2} + \frac{\pi}{L}\left(\left(\hat\rho^+(0) + eY\right)^2 + \left(\hat\rho^-(0) - eY\right)^2 + \left(\hat\rho^+(0) + eY\right) 2 W_0 \left(\hat\rho^-(0) - eY\right) \right) + \nonumber \\ \nopagebreak \hat{\int}_{\dLam\backslash \{0\}}\hat{\rm d} k\, \xx \left( \frac{e^2}{4\pi k^2} \hat\rho(k)\hat\rho(-k) + \frac{1}{2}\left( \hat\rho^+(-k)\hat\rho^+(k) + \hat\rho^-(k)\hat\rho^-(-k) \right) + \hat\rho^+(k)W_k \hat\rho^-(-k)\right)\xx. \end{eqnarray} \subsection{Diagonalization of the Hamiltonian} \label{zeromode} {}Following \cite{HSU} we now write \begin{equation} H = \frac{2\pi}{L}\sum_{k\geq 0} h_k. \end{equation} Introducing boson creation-- and annihilation operators \alpheqn \begin{equation} \label{crho} c(k) = \left\{\bma{cc} \frac{1}{\sqrt{|k|}}\hat\rho^+(k) & \mbox{ for $k>0$}\\ \frac{1}{\sqrt{|k|}}\hat\rho^-(k) & \mbox{ for $k<0$} \end{array}\right. \end{equation} obeying usual CCR \begin{equation} \ccr{c(k)}{c^*(p)} = \hat{\delta}(k-p)\quad \mbox{etc.}. \end{equation} We then get for $h_{k>0}$ \begin{equation} h_k =\left(k + \frac{e^2}{2\pi k}\right)\left(c^*(k)c(k)+c^*(-k)c(-k) \right) + \left( kW_k + \frac{e^2}{2\pi k}\right)\left(c^*(k)c^*(-k)+c(k)c(-k) \right). \end{equation} \reseteqn {}For $k=0$ we introduce the quantum mechanical variables \alpheqn \begin{eqnarray} \label{qmv} P &=& \left( \hat\rho^+(0) -\hat\rho^-(0) + 2eY \right)\, , \nonumber \\ \nopagebreak X &=& {\rm i}\frac{L}{2\pi}\frac{1}{2e} \frac{\partial}{\partial Y} \end{eqnarray} obeying Heisenberg relations, $\ccr{P}{X}=-{\rm i} L/2\pi$, which allow us to write $h_0$ as Hamiltonian of a harmonic oscillator, \begin{equation} \label{zeromom0} h_0= \frac{e^2}{\pi} X^2 + \frac{1}{4}(1-W_0)\, P^2 + \frac{1}{4} (1+W_0)\, Q_0^2 -\frac{1}{2}\sqrt{\frac{e^2}{\pi}}\frac{L}{2\pi} \end{equation} \reseteqn (the last term stems from normal ordering $\,\xx\cdots\xx\,$ and is irrelevant for the following). We can now solve the model by diagonalizing its decoupled Fourier modes $h_k$ separately, with the help of a boson Bogoliubov transformation preserving the CCR, \begin{equation} \label{BT} C(k) = \cosh(\lambda_k) c(k) + \sinh(\lambda_k) c^*(-k) \end{equation} where $\lambda_k=\lambda_{-k}$. This leads to \alpheqn \begin{equation} h_k = \omega_k\left(C^*(k)C(k) + C^*(-k)C(-k)\right) -2\eta_k \frac{L}{2\pi} \end{equation} if we choose \begin{equation} \tanh(2\lambda_k) = \frac{2\pi k^2 W_k + e^2}{2\pi k^2 + e^2} \label{th}\, . \end{equation} Then \begin{equation} \omega_k^2 = k^2(1-W_k^2) + \frac{e^2}{\pi}(1-W_k) \end{equation} and \begin{equation} \eta_k = \frac{1}{2}\left(|k| + \frac{e^2}{2\pi|k|} - \omega_k \right) \quad (k\neq 0) . \end{equation} \reseteqn The zero--momentum piece $h_0$ is just a harmonic oscillator and can be written as \alpheqn \begin{equation} \label{zeromom} h_0 = \omega_0 C^*(0) C(0) + \frac{1}{4} (1+W_0)\, Q_0^2 - \eta_0\frac{L}{2\pi} \end{equation} with \begin{equation} \label{Cnull} C(0) = \frac{1}{\sqrt 2}\left(rX + \frac{1}{r}{\rm i} P\right),\quad r^4=\frac{e^2}{\pi}\frac{4}{1-W_0}, \end{equation} energy--squared \begin{equation} \omega_0^2 = \frac{e^2}{\pi}(1- W_0) \end{equation} and zero point energy \begin{equation} \eta_0 =\frac{1}{2}\left(\sqrt{\frac{e^2}{\pi}} - \sqrt{\frac{e^2}{\pi}(1-W_0)}\, \right). \end{equation} \reseteqn Thus we get the Hamiltonian in the following form \alpheqn \begin{equation} H=\hat{\int}_{\dLam}\hat{\rm d} k \omega_k C^*(k)C(k) -L E_0 \end{equation} with the ground state energy density given by \begin{equation} E_0=\frac{1}{2\pi}\hat{\int}_{\dLam}\hat{\rm d} k\, \eta_k. \end{equation} \reseteqn (Note that for large $|k|$, $ \eta_k = \frac{1}{2}\left( \frac{1}{2}|k W_k^2| + \frac{e^2}{2\pi |k|}W_k\right)\left(1+{\cal O}\left(\frac{1}{|k|}\right)\right), $ hence $E_0$ is finite due to our assumptions \Ref{condition1} on the potential.) We now construct the unitary operator ${\cal U}$ implementing the Bogoliubov transformation \Ref{BT}, i.e.\ \begin{equation} C(k) = {\cal U} c(k){\cal U}^* \quad \forall k\in\Lambda^* . \end{equation} It is easy to see that operators ${\cal U}_k$ satisfying $C(\pm k) = {\cal U}_k c(\pm k){\cal U}_k^*$ for all $k>0$ are given by \alpheqn \begin{equation} \label{cU} {\cal U}_k = \ee{S_k},\quad S_k = \lambda_k \left( c(k)c(-k) - c^*(k)c^*(-k) \right) \end{equation} which are unitary since the operators $S_k$ are screw-hermitian.\footnote{i.e.\ ${\rm i} S_k$ is selfadjoint} Thus, \begin{equation} {\cal U} = \ee{S},\quad S=\sum_{k>0}S_k . \end{equation} This operator $S$ can be shown to exist and defines an anti-selfadjoint operator if and only if \begin{equation} \sum_{k>0}|k||\lambda_k|^2<\infty \label{24c}, \end{equation}\reseteqn and therefore (\ref{24c}) is necessary and sufficient for the unitary operator ${\cal U}$ to exist. This latter condition is equivalent to the second one in \Ref{condition1} and thus fulfilled by assumption. Note that \begin{equation} \label{hdiag} {\cal U}^* H {\cal U} = \frac{2\pi}{L}h_0 + \frac{2\pi}{L}\sum_{k>0} \left( \omega_k\left( c^*(k)c(k) + c^*(-k)c(-k)\right) -2\eta_k\frac{L}{2\pi}\right) \equiv H_D \end{equation} and therefore ${\cal U}$ is the unitary operator diagonalizing the non--zero modes of our Hamiltonian. \subsection{Gauge invariant states} \label{GND} By the gauge fixing above we reduced the Hilbert space from ${\cal H}$ to ${\cal H}'_{{\rm phys}}$ containing all states invariant under {\em small} gauge transformations, i.e. of the form $\ee{{\rm i}\alpha(x)}$ with $\alpha(L/2)=\alpha(-L/2)$. There are, however, still large gauge transformations present which are generated by $\ee{\ii2\pi x/L}$. It is important to note that physical states need not be invariant under these latter transformations, but it is useful to construct states with simple transformation properties. This is the origin of the $\theta$--vacuum. The large gauge transformation $\ee{\ii2\pi x/L}$ acts on the fields as follows \begin{eqnarray} \label{large} \psi(x)&\stackrel{R}{\to}& \ee{{\rm i} 2\pi x/L}\psi(x) = (R_+R_-)^{-1} \psi(x)(R_+R_-)\, ,\nonumber \\ \nopagebreak eY&\stackrel{R}{\to}& eY - 1 \end{eqnarray} where $R_\pm$ are the implementers of $\ee{{\rm i} 2\pi x/L}$ in the chiral sectors of the fermions and are discussed in detail in Appendix A. The large gauge transformation $R$ obviously generates a group ${\bf Z}$, $n\to R^n$, and we denote this group as ${\bf Z}_R$. Our aim is to construct the states in ${\cal H}_{{\rm phys}}$ which carry an irreducible representation of ${\bf Z}_R$ and especially the ground states of our model. We start with recalling that the Fermion Fock space can be decomposed in sectors of different chiral charges $\hat\rho^\pm(0)$, $${\cal H}_{{\rm Fermion}} =\bigoplus_{n_+,n_-\in{\bf Z}}{\cal H}^{(n_+,n_-)}$$ where $$ {\cal H}^{(n_+,n_-)}=\left\{\left. \Psi\in {\cal H}_{{\rm Fermion}}\right| \hat\rho^\pm(0)\Psi = n_\pm \Psi \right\} = R_+^{n_+}R_-^{-n_-}{\cal H}^{(0,0)} $$ (for a more detailed discussion see Appendix A). Thus, \begin{equation} {\cal H}_{{\rm phys}}= {\cal L}^2({\bf R},{\rm d} Y)\otimes {\cal H}_{{\rm Fermion}}' \end{equation} where \begin{equation} {\cal H}_{{\rm Fermion}}' = \bigoplus_{n\in{\bf Z}}{\cal H}^{(n,-n)}, \quad {\cal H}^{(n,-n)} = (R_+ R_-)^n{\cal H}^{(0,0)} \end{equation} is the zero charge subspace of the Fermion Fock space and we use the Schr\"odinger representation for the physical degree of freedom $Y = \int_{\Lam} {\rm d}{x} A_1(x)/2\pi$ of the photon field as discussed in the last subsection. ${\cal H}_{{\rm phys}}$ can therefore be spanned by states \alpheqn \begin{equation} \label{tetn} \Psi(n) = \phi\mbox{$(Y+\frac{n}{e}) $} (R_+R_-)^n \Psi,\quad \phi \in {\cal L}^2({\bf R},{\rm d} Y),\, \Psi\in{\cal H}^{(0,0)} \end{equation} which, under a large gauge transformation \Ref{large}, transform as \begin{equation} \Psi(n)\stackrel{R}{\to} \Psi(n-1). \end{equation} \reseteqn Thus the states transforming under an irreducible representation of ${\bf Z}_R$ are given by \begin{equation} \label{tet} \Psi^\theta= \sum_{n\in{\bf Z}}\ee{{\rm i}\theta n}\Psi(n) \stackrel{R}{\to} \ee{{\rm i}\theta}\Psi^\theta \end{equation} It is easy to calculate the inner products of these states, \begin{equation} <\Psi_1^\theta,\Psi_2^{\theta'}>= 2\pi\delta_{2\pi}(\theta-\theta') <\Psi_1,\Psi_2>_{{\rm F}}(\phi_1,\phi_2)_{{\cal L}^2} \end{equation} ($2\pi\delta_{2\pi}(\theta)=\sum_{n\in{\bf Z}}\ee{{\rm i} n\theta}$, since $<(R_+R_-)^n\Psi_1,(R_+R_-)^m\Psi_2> = \delta_{n,m}<\Psi_1,\Psi_2>_{{\rm F}}$; $<\cdot,\cdot>_{{\rm F}}$ and $<\cdot,\cdot>_{{\cal L}^2}$ are the inner products in ${\cal H}_{{\rm Fermion}}$ and ${\cal L}^2({\bf R},{\rm d} Y)$, respectively). Thus the states $\Psi^\theta$ actually are not elements in ${\cal H}_{{\rm phys}}$ (they do not have a finite norm). In our calculation of Green functions below we find it useful to use the notation \begin{equation} \label{reg} <\Psi_1^\theta,\Psi_2^{\theta}>_\theta \equiv <\Psi_1,\Psi_2>_{{\rm F}}(\phi_1,\phi_2)_{{\cal L}^2} \end{equation} which can be regarded as redefinition of the inner product using a simple multiplicative regularization (dropping the infinite term $2\pi\delta_{2\pi}(0)$). We now construct the ground states of our model. As expected, the quantum mechanical variables $P,X$ \Ref{qmv} describing the zero mode $h_0$ of the Hamiltonian have a simple representation on the $\theta$-states \Ref{tet}, \begin{eqnarray} P\Psi^\theta = \sum_{n\in{\bf Z}}\ee{{\rm i}\theta n} 2e\mbox{$(Y+\frac{n}{e})$} \phi\mbox{$(Y+\frac{n}{e}) $}(R_+R_-)^n\Psi \, ,\nonumber \\ \nopagebreak X\Psi^\theta = \sum_{n\in{\bf Z}}\ee{{\rm i}\theta n} \frac{{\rm i}}{2e}\frac{L}{2\pi}\frac{\partial}{\partial Y} \phi\mbox{$(Y+\frac{n}{e}) $} (R_+R_-)^n\Psi. \end{eqnarray} Thus the ground states of $h_0$ annihilated by $C(0)$ are of the form \Ref{tetn} with \begin{equation} \label{gst} \phi_0(Y) = \left(\frac{\pi}{4e^2\alpha}\right)^\frac{1}{4} \exp\left(-\alpha(2eY)^2\right) \end{equation} where $ \label{alpha} \alpha=\frac{1}{L}\sqrt{\frac{\pi^3}{2e^2}(1-W_0)}, $ and the other eigenstates are the harmonic oscillator eigenfunctions $\phi_n\propto C^*(0)^n\phi_0$. From $C(k)={\cal U} c(k){\cal U}^*$ and $c(k)\Omega_{{\rm F}}=0$ it is clear that the ground state of all $h_{k>0}$ is ${\cal U} \Omega_{{\rm F}}$. We conclude that the ground states of our model obeying $H \Psi_0^\theta = L E_0\Psi_0^\theta$ are given by \begin{equation} \label{vac} \Psi_0^\theta= \sum_{n\in{\bf Z}}\ee{{\rm i}\theta n} \phi_0\mbox{$(Y+\frac{n}{e}) $}(R_+R_-)^n {\cal U}\Omega_{{\rm F}}. \end{equation} \subsection{Gauge invariant Green functions} \label{green} The observables of our model now are operators on ${\cal H}_{\rm phys}$ where $\int_\Lambda {\rm d} x A_1(x)$ is represented by $2\pi Y$. We recall that the fully gauge invariant field operators are the $\chi$, \Ref{chi}, which are represented in the present gauge fixed setting by $$\chi_\sigma(x)=\ee{{\rm i} 2\pi e Y(x-r)/L}\psi_\sigma(x).$$ These operators depend on the $r\in\Lambda$ chosen. Bilinears such as meson operators are, however, independent of $r$ and give rise to translational invariant equal time Green functions. Moreover, on the quantum level not only the Wilson line $W[A_1]$ \Ref{Wilson} but actually even \begin{equation} e \int_\Lambda{\rm d} x A_1(x) + \mbox{$\frac{1}{2}$} Q_5 \equiv w[A_1] \end{equation} is gauge invariant (note that $W[A_1] = \ee{{\rm i} w[A_1]}$). This operator is represented by $e Y+\mbox{$\frac{1}{2}$} Q_5 = P/2$ (cf. \Ref{qmv}). The gauge invariant equal time Green functions of the model are the ground state expectation values of products $(\cdots)$ of meson operators and functionals $F[P,X]$ of the zero mode operators $P$, $X$. Since we only consider $(\cdots)$ which are also invariant under large gauge transformations, the transition amplitudes $\left<\Psi_1^\theta,(\cdots)\Psi_2^{\theta'} \right>$ are always proportional to $2\pi\delta(\theta-\theta')$. Thus the Green functions we consider can be defined as \begin{equation} \label{greenf} \left< \Psi_0^\theta, F[P,X] \, \chi^*_{\sigma_1}(x_1)\chi_{\tau_1}(y_1) \cdots \chi^*_{\sigma_N}(x_N) \chi_{\tau_N}(y_N)\Psi_0^\theta \right>_\theta \end{equation} (note that $\left< \Psi_0^\theta, \Psi_0^\theta \right>_\theta =1$, cf. \Ref{reg}). {}Following \cite{HSU} it is useful to define {\em interacting fermion fields} \alpheqn \begin{equation} \label{intf} \Psi_\sigma(x) = {\cal U}^* \psi_\sigma(x)\, {\cal U} \end{equation} such that (\ref{greenf}) becomes \begin{equation} \mbox{Eq.\ \Ref{greenf}} = \left<\Omega^\theta, F[P,X]\, \Psi^*_{\sigma_1}(x_1) \Psi_{\tau_1}(y_1)\cdots \Psi^*_{\sigma_N}(x_N) \Psi_{\tau_N}(y_N) \Omega^{\theta'}\right> \end{equation} where \begin{equation} \label{freevac} \Omega^\theta= \sum_{n\in{\bf Z}}\ee{{\rm i}\theta n} \phi_0\mbox{$(Y+\frac{n}{e}) $}(R_+R_-)^n \Omega_{{\rm F}} \end{equation} \reseteqn is the $\theta$--state constructed from the free fermion vacuum. The strategy to calculate Green functions of the model using bosonization techniques is the following: the relation \Ref{k2} of appendix A can be used to move the operators $R_\pm$ and combine them to some power of $(R_+R_-)$. The operators $Q_\pm$ when applied to physical states become simple ${\bf C}$--numbers: $Q_\pm (R_+R_-)^n = (R_+R_-)^n (\pm n+Q_\pm)$ for all integers $n$, and $Q_\pm\Omega_{\rm F} = 0$. For the exponentials of boson operators we use the decomposition into creation and annihilation parts outlined in A.4. The normal ordering procedure gives a product of exponentials of commutators which are (${\bf C}$-number) functions. For the correlation functions of meson operators $\chi_\sigma^*(x)\chi_{\sigma'}(y)$ we obtain: \alpheqn \begin{eqnarray} \left<\Psi_0^\theta,\chi^*_{\pm}(x) \chi_{\pm}(y)\Psi_0^{\theta} \right>_\theta &=& \ee{-\frac{\pi}{4L}m(x-y)^2} \ee{\Delta(x-y)}g_0^\pm(x-y) \label{2p1}\, , \\ \left<\Psi_0^\theta,\chi_{\pm}^*(x) \chi_{\mp}(y)\Psi_0^{\theta} \right>_\theta &=& \ee{\mp {\rm i}\theta} \ee{-i\frac{2\pi}{L}(x-y)} \ee{-\frac{\pi m}{4L}((x-y)+\frac{2}{m})^2} C(L) \ee{D(x-y)} \label{2p2} \, . \end{eqnarray} \reseteqn with \begin{eqnarray} \Delta &=& \sum_{k>0}\frac{2\pi}{Lk}\sinh^2(\lambda_k)[\ee{{\rm i} kx }+ \ee{-{\rm i} kx }-2] \nonumber\, , \\ D(x) &=& -\sum_{k>0} \frac{\pi}{Lk} \sinh(2\lambda_k) [ \ee{{\rm i} kx }+ \ee{-{\rm i} kx }-2] \label{DeDC}\, , \\ C(L) &=& \frac{1}{L}\exp[\sum_{k>0} \frac{2\pi}{kL}(\sinh(2\lambda_k)-2\sinh^2(\lambda_k))]\,. \nonumber \end{eqnarray} where $g_0^\pm(x)=\frac{1}{L}\frac{e^{\mp i\frac{\pi}{L}x}}{1-e^{\pm i\frac{2\pi}{L} (x\pm i\varepsilon)}}$ is the 2-point function of free fermions, and the Schwinger mass is renormalized to $m^2=\frac{e^2}{\pi(1-W_0)}$. Note that the Green function \Ref{2p2} depends on $\theta$ and is non--zero due to chiral symmetry breaking as in the Schwinger model. As expected, for vanishing electromagnetic coupling, $e=0$, this Green function vanishes (due to the factor $\ee{-\pi/mL}$ appearing in (\ref{2p2})). {}From (\ref{2p2}) we can calculate the chiral condensate by setting $x=y$, and in the limit $L\to\infty$ we obtain \begin{equation} \lim_{L\to\infty} \left<\Psi_0^\theta,\chi_{\pm}^*(x) \chi_{\mp}(x)\Psi_0^{\theta} \right>_\theta = \lim_{L\to\infty} \ee{\mp {\rm i}\theta} \ee{-\frac{\pi}{mL}} C(L) \nonumber = \ee{\mp {\rm i}\theta} C \end{equation} with a constant $C$ which can be calculated in principle from eq.\ (\ref{DeDC}). In the special case of the Schwinger model ($W_k=0$), $C$ can be computed and we recover the well--known result $C_{W_k=0}=\frac{m}{4\pi}\ee{\gamma}$ where $\gamma=0.577\ldots$ is Eulers constant (see e.g.\ \cite{SaWi,Hoso}). \section{Multiplicative regularization and the Thirring-Schwinger model} We recall that the Thirring model is formally obtained from the Luttinger model in the limits \begin{equation} \label{LT} L\to \infty, \quad V(x)\to g \delta(x) \end{equation} i.e.\ when the interaction becomes local and space becomes infinite. The first limit amounts to remove the IR cut--off of our model. By inspection it can be easily done in all Green functions. The second limit in \Ref{LT} is non--trivial: we recall, that condition \Ref{condition1} on the Luttinger potential requires sufficient decay of the Fourier modes $W_k$ of the interaction, and this is violated in the Thirring model where $W_k=W_0$ is independent of $k$. This latter condition was necessary for the interacting model to be well--defined on the Hilbert space of the non--interacting model. A better understanding can be obtained by explicitly performing the limit \Ref{LT} in the present setting. The idea is to find a family of Luttinger potentials $\{ V_\ell(x)\}_{\ell>0}$ becoming local for $\ell\downarrow 0$, i.e.\ for all $\ell>0$ the condition \Ref{condition1} is fulfilled and $\lim_{\ell\downarrow 0} V_\ell(x)=g\delta(x)$. Then for all $\ell>0$ everything is well-defined on the free Hilbert space and one can work out in detail how to regularize such that the correlation functions make sense for $\ell\downarrow 0$. We note that a direct construction of the Thirring model in a framework similar to the one here has been completed in \cite{CRW}. This construction seems to be, however, different from the one outlined below. {}For the case of Luttinger-Schwinger model we split the function $\Delta(x)$ into a part corresponding to the pure Luttinger model and a part which describes the additional Schwinger coupling, i.~e.~$\Delta(x)=(\Delta(x)-\Delta^{e=0}(x))+\Delta^{e=0}(x)$. The limit $W_k=W_0={\rm const.}$ exists for $\Delta-\Delta^{e=0}$. As $L\to\infty$, the sum in (\ref{DeDC}) turns into an integral and we obtain \begin{eqnarray} \Delta(x)-\Delta^{e=0}(x)=&&\!\!\!\!\!\!\! \frac{1}{\sqrt{1-W_0^2}} \int_0^\infty {\rm d}k\, \left( \frac{1}{\sqrt{k^2+\mu^2}}-1\right)(\cos (kx) -1)+ \nonumber\\ && \!\!\!\!\!\!\! \sqrt{\frac{1+W_0}{1-W_0}}\, \int_0^\infty {\rm d}k\, \frac{\mu^2}{k^2}\frac{1}{\sqrt{k^2+\mu^2}} (\cos (kx) -1) \, . \end{eqnarray} The first integral becomes $K_0(|\mu x|)+\ln\frac{|\mu x|}{2}+\gamma$ and the expression in the last line is a second integral $(n=2)$ of $K_0$ defined iteratively by Ki$_n(x)=\int_x^\infty $Ki$_{n-1}(t) \, dt$, Ki$_0=K_0$ \cite{Abram}. Moreover we introduced a new mass by $\mu^2=e^2/(\pi(1+W_0))$. Note that the singularities at the origin of the Bessel function are removed by the additional terms, consistent with $\Delta(0)=0$. No regularization has been necessary so far. Renormalization comes along with $\Delta^{e=0}$. We choose a Luttinger-interaction such that $(1-W_k^2)^{-1/2}-1=2a^2 e^{-\ell k}$ where $\ell$ defines the range of the interaction. For this choice we find \begin{eqnarray} \Delta^{e=0}(x)=2a^2\ln\left|\frac{\ell}{x+i\ell}\right| \end{eqnarray} and obviously the Thirring limit makes sense only if one removes the singular part $\ln \ell$ which can be done by a wave function renormalization of the form \begin{eqnarray} \chi_\pm(x) \to \tilde\chi_\pm(x)=Z^{1/2}(a,\ell)\chi_\pm(x) \quad \makebox{with} \quad Z^{1/2}(a,\ell)=\ell^{-a^2} \, . \label{Th3} \end{eqnarray} A similar discussion holds for the chirality mixing correlation function. The 2-point function of the Thirring-Schwinger model therefore become \alpheqn \begin{eqnarray} \langle\Psi_0^\theta,\tilde\chi_\pm^*(x)\tilde\chi_\pm(0)\Psi_0^\theta\rangle_\theta &=&e^{\Delta_{\rm reg}(x)}g_0^\pm(x)\, ,\\ \left<\Psi_0^\theta,\tilde\chi_{\pm}^*(x) \tilde\chi_{\mp}(0)\Psi_0^{\theta} \right>_\theta &=& \ee{\mp {\rm i}\theta} C_{\rm reg}\, \ee{D_{\rm reg}(x)} \, . \end{eqnarray} \reseteqn If we define $\tau_0$ by $\tanh(2\tau_0)=W_0$ we can write \begin{eqnarray} \Delta_{\rm reg}(x)&\!=&\!\cosh (2\tau_0)\left[ K_0(|\mu x|)+\ln\frac{|\mu x|}{2}+\gamma\right] +\nonumber\\ &&\!\!\frac{1}{2}e^{2\tau_0}\left[1-\frac{\pi}{2}|\mu x|-{\rm Ki}_2(|\mu x|)\right]+ (\cosh (2\tau_0)-1)\ln |x| \, , \nonumber\\ D_{\rm reg}(x) &\!=&\! -\sinh (2\tau_0)\left[ K_0(|\mu x|)+\ln\frac{|\mu x|}{2}+\gamma\right] - \\ &&\!\! \frac{1}{2}e^{2\tau_0}\left[1-\frac{\pi}{2}|\mu x|-{\rm Ki}_2(|\mu x|) \right] \, ,\nonumber\\ \ln C_{\rm reg} &\!=&\! \gamma+\ln\frac{1}{2\pi}+e^{-2\tau_0}\ln \frac{\mu}{2} \nonumber \, . \end{eqnarray} We checked that all Green functions of the Thirring-Schwinger model have a well-defined limit after the wave function renormalization. We would like to stress that this procedure can be naturally interpreted as low--energy limit of the Luttinger--Schwinger model: if one is interested only in Green functions describing correlations of far--apart fermions, the precise form of the Luttinger interaction $V(x)$ should be irrelevant and only the total interaction strength $g = \int{\rm d} x\, V(x)$ should matter. Thus as far as these correlators are concerned, they should be equal to the ones of the Thirring model corresponding to this coupling $g$. \section{Conclusion} We formulated and solved the Luttinger-Schwinger model in the Hamiltonian formalism. Structural issues like gauge invariance, the role of anomalies and the structure of the physical states were discussed in detail. The necessary tools for computing all equal time correlation functions were prepared and illustrated by calculating the 2--point Green functions. From this the chiral condensate and critical exponents were computed. We could also clarify how the non trivial short distance behavior of the Thirring-Schwinger model arises in a limit from the Luttinger-Schwinger model. \app \section*{Appendix A: Bosons from fermions and vice versa} In this appendix we summarize the basics for the bosonization used in the main text to solve the Luttinger--Schwinger model. Bosonization is known in the physics literature since quite some time (\cite{CR,PressSegal,Kac,Mickelsson}), for a discussion of the older history see \cite{HSU}). We consider the fermion Fock space ${\cal H}_{\rm Fermion}$ generated by the fermion field operators from the vacuum $\Omega_{\rm F}$ as described in the main text. We note that ${\cal H}_{\rm Fermion}={\cal H}_{+}\otimes{\cal H}_{-}$ where ${\cal H}_{\pm}$ are generated by the left-- and right--handed chiral components $\hat\psi_+$ and $\hat\psi_-$ of our Dirac fermions. Bosonization can be formulated for the chiral components $\hat\psi_\pm$ separately as it leaves the two chiral sectors ${\cal H}_{\pm}$ completely decoupled. For our purpose it is more convenient to treat both chiral sectors together. \subsection*{A.1 Structure of fermion Fock space} We start by introducing two unitary operators $R_\pm$ which are defined up to an irrelevant phase factor (which we will leave unspecified) by the following equations, \begin{equation} \hat\psi_\pm(k)R_\pm = R_\pm \hat\psi_\pm (k - \frac{2\pi}{L}) \end{equation} and $R_\pm$ commutes with $\hat\psi_\mp$. A proof of existence and an explicit construction of these operators can be found in \cite{Ruij}. Here we just summarize their physical meaning and special properties. It is easy to see that $R_\pm$ are just the implementors of Bogoliubov transformations given by the {\em large gauge transformations} $\psi_\pm(x)\mapsto \ee{{\rm i} 2\pi x/L}\psi_\pm(x)$ and $\psi_\mp(x)\mapsto \psi_\mp(x)$, hence $R_+R_-$ and $R_+R_-^{-1}$ implement the vector-- and the axial large gauge transformations $\ee{{\rm i} 2\pi x/L}$ and $\ee{{\rm i}\gamma_5 2\pi x/L }$, respectively. These have non--trivial winding number\footnote{ the w.n. of a smooth gauge transformation $\Lambda\to{\rm U}(1)$, $x\mapsto \ee{{\rm i}\alpha(x)}$ is the integer $\frac{1}{2\pi}\left(\alpha(L)-\alpha(0)\right)$. } and change the vacuum $\Omega_{\rm F}$ to states containing (anti-) particles. The latter follows from the commutator relations with the chiral fermion currents \begin{eqnarray} (R_\pm)^{-1} \hat\rho^\pm (k) R_\pm = \hat\rho^\pm (k) \pm \delta_{k,0} \label{k2}\, . \end{eqnarray} The essential point of bosonization is that the total Hilbert space ${\cal H}_{\rm Fermion}$ can be generated from $\Omega_{\rm F}$ by the chiral fermion currents $\hat\rho^{\pm}(k)$ and $R_\pm$. More precisely, for all pairs of integers $n_+,n_-\in{\bf Z}$ we introduce the subspaces ${\cal D}^{(n_+,n_-)}$ of ${\cal H}_{{\rm Fermion}}$ containing all linear combinations of vectors \begin{equation} \label{lin} \hat\rho^+(k_1) \cdots \hat\rho^+(k_{m_+}) \hat\rho^-(q_1)\cdots \hat\rho^-(k_{m_-} ) R_+^{n_+} R_-^{-n_-} \Omega_F \end{equation} where $m_\pm\in{\bf N}_0$ and $k_i,q_i\in\Lambda^*$. The basic result of the boson--fermion correspondence is the following {\bf Lemma:} The space \begin{equation} {\cal D} \equiv \bigoplus_{n_+,n_-\in{\bf Z}} D^{(n_+,n_-)} . \end{equation} is dense in ${\cal H}_{\rm Fermion}$ (for a proof see e.g.\ \cite{CR}). {\em Remark:} This Lemma gives the following picture of the structure of the Fock space ${\cal H}_{\rm Fermion}$: It splits into {\em superselection sectors} ${\cal H}^{(n_+,n_-)}$ (which are the closure of ${\cal D}^{(n_+,n_-)}$) containing the eigenstates of the chiral charges $Q_\pm$ with eigenvalues $n_\pm$. The fermion currents $\hat\rho^\pm(k)$ leave all these sectors invariant, and the operators $R_\pm$ intertwine different sectors, $R_+: {\cal H}^{(n_+,n_-)}\to {\cal H}^{(n_+ +1,n_-)}$ and $R_-: {\cal H}^{(n_+,n_-)}\to {\cal H}^{(n_+,n_- - 1)}$. \subsection*{A.2 Kronig's identity} The basic formula underlying the solution of our model is \begin{equation} \label{kronig} \hat H_0 = \frac{\pi}{L}\left( Q_+^2 + Q_-^2 \right) + \frac{2\pi}{L} \sum_{k>0}\left( \hat\rho^+(-k)\hat\rho^+(k) + \hat\rho^-(k) \hat\rho^-(-k) \right) . \end{equation} It expresses the free Dirac Hamiltonian in terms of bilinears of the chiral fermion currents. \subsection*{A.3 Boson--fermion correspondence} The boson--fermion correspondence provides explicit formulas of the fermion operators $\psi_\pm(x)$ in terms of operators $\hat\rho^\pm(k)$ and $R_\pm$, \aalpheqn \begin{equation} \label{limit} \psi_\pm(x) = \lim_{\varepsilon\searrow 0} \psi_\pm(x;\varepsilon) \end{equation} (this limit can e.g. be understood in the weak sense for states in ${\cal D}$), where \begin{eqnarray} \psi_\pm(x;\varepsilon) = \frac{1}{\sqrt{L}} S_\pm(x) \normal{\exp(K_\pm(x;\varepsilon))} \label{bfc} \end{eqnarray} with \begin{equation} \label{S} S_\pm(x)= \ee{\pm {\rm i} \pi x Q_\pm/L } (R_\pm)^{\mp 1}\ee{\pm {\rm i} \pi x Q_\pm/L } = \ee{\mp \pi{\rm i} x/L} (R_\pm)^{\mp 1}\ee{\pm {\rm i} 2\pi x Q_\pm/L } \end{equation} and \begin{equation} \label{Kpm} K_\pm(x;\varepsilon) =\mp \frac{2\pi}{L} \sum_{k\in\Lambda^*\backslash\{0\}} \frac{\hat\rho^\pm(-k)}{k}\ee{-{\rm i} kx}\ee{-\varepsilon|k|} = - K_\pm(x;\varepsilon)^*. \end{equation} \areseteqn More explicitly, the normal ordering $\normal{\cdots}$ is with respect to the fermion vacuum $\Omega_{{\rm F}}$ (cf.\ \Ref{vacF}), \aalpheqn \begin{equation} \normal{\exp(K_\pm(x;\varepsilon))} = \exp(K^{(-)}_\pm(x;\varepsilon)) \exp(K^{(+)}_\pm(x;\varepsilon)) \end{equation} where \begin{equation} K^{(\sigma)}_\pm(x;\varepsilon) = \sigma\frac{2\pi}{L}\sum_{k>0} \frac{\hat\rho^\pm(\pm \sigma k)}{k}\ee{\pm\sigma {\rm i} kx}\ee{-\varepsilon|k|}, \quad \sigma=+,- \end{equation} \areseteqn is such that $K_\pm = K_\pm^{(-)}+K_\pm^{(+)}$ and $K_\pm^{(+)}\Omega_{{\rm F}} = 0$ (cf.\ \Ref{vacF}). \subsection*{A.4 Interacting fermions} {}From the definition of the interacting fermion fields $\Psi(x)$ (\ref{intf}) and the representation of free fermions in terms of bosons, we are led to investigate the interacting kernel $\tilde K_\pm(x)={\cal U}^* K_\pm(x) {\cal U}$: $$ \tilde{K}_\pm(x) =\mp \frac{2\pi}{L} \sum_{k\in\Lambda^*\backslash\{0\}} \frac{1}{k} \left( \cosh(\lambda_k)\, \hat\rho^\pm(-k) - \sinh(\lambda_k)\, \hat\rho^\mp(-k) \right) \ee{-{\rm i} kx}\ee{-\varepsilon|k|}\, . $$ It is convenient to write \newcommand{{K\! s}}{{K\! s}} \newcommand{{K\! c}}{{K\! c}} \newcommand{{K\! s/c}}{{K\! s/c}} \aalpheqn \begin{equation} \tilde{K}_\pm = \tilde{K}_\pm^{(+)} + \tilde K_\pm^{(-)},\quad \tilde{K}_\pm^{(\sigma)} = Kc_\pm^{(\sigma)} -Ks_\mp^{(\sigma)} \end{equation} where the upper index refers to the creation-- ($\sigma=-$) and annihilation-- ($\sigma=+$) parts of operators and \begin{eqnarray} {K\! c}^{(\sigma)}_\pm(x) &=& \sigma\sum_{k>0}\frac{2\pi}{Lk}\cosh(\lambda_k) \hat\rho^\pm(\pm\sigma k)\ee{\mp\sigma {\rm i} kx }\ee{-\varepsilon k} \, , \nonumber \\ \nopagebreak {K\! s}^{(\sigma)}_\pm(x) &=& \sigma\sum_{k>0}\frac{2\pi}{Lk}\sinh(\lambda_k) \hat\rho^\pm(\pm\sigma k)\ee{\mp\sigma {\rm i} kx }\ee{-\varepsilon k} \: . \end{eqnarray} \areseteqn The nonzero commutators of these operators are \begin{eqnarray} \ccr{{K\! c}^{(+)}_\pm(x)}{{K\! c}^{(-)}_\pm(y)} &=& - \sum_{k>0}\frac{2\pi}{Lk}\cosh^2(\lambda_k) \ee{ \mp {\rm i} k(x-y) }\ee{-2\varepsilon k} \, , \nonumber \\ \nopagebreak \ccr{{K\! c}^{(+)}_\pm(x)}{{K\! s}^{(-)}_\pm(y)} &=& - \sum_{k>0}\frac{\pi}{Lk}\sinh(2\lambda_k) \ee{ \mp{\rm i} k(x-y) }\ee{-2\varepsilon k}\, , \\ \ccr{{K\! s}^{(+)}_\pm(x)}{{K\! s}^{(-)}_\pm(y)} &=& - \sum_{k>0}\frac{2\pi}{Lk}\sinh^2(\lambda_k) \ee{ \mp{\rm i} k(x-y) }\ee{-2\varepsilon k} \: . \nonumber \end{eqnarray} We find the following normal ordered expression for the interacting fermions \begin{equation} \Psi_\pm(x) = \frac{1}{\sqrt{L}} z S_\pm(x) \normal{\ee{\tilde K_\pm(x) }} \end{equation} where $z = \ee{- \sum_{k>0} \frac{2\pi}{Lk}\sinh^2(\lambda_k) }$. \appende \begin{center}{\bf Acknowledgments}\end{center} E.L. would like to thank the Erwin Schr\"odinger International Institute in Vienna for hospitality where part of this work was done, and the ``\"Osterreichische Forschungsgemeinschaft'' for partial financial support in May/June 1994 when this work was begun. He would also like to thank S.G. Rajeev and Mats Wallin for usefull discussions. The authors thank the referee for valuable suggestions concerning the presentation of their results.
proofpile-arXiv_065-408
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} This paper is focused on the actual problem of heavy flavor physics, exclusive s.l. decays of low-lying bottom and charm baryons. Recently, activity in this field has obtained a great interest due to the experiments worked out by the CLEO Collaboration \cite{CLEO} on the observation of the heavy-to-light s.l. decay $\Lambda_c^+\to\Lambda e^+\nu_e$. Also the ALEPH~\cite{ALEPH} and OPAL~\cite{OPAL} Collaborations expect in the near future to observe the exclusive mode $\Lambda_b\to\Lambda_c\ell\nu$. In ref. \cite{Aniv} a model for QCD bound states composed from light and heavy quarks was proposed. Actually, this model is the Lagrangian formulation of the NJL model with separable interaction \cite{Goldman} but its advantage consists in the possibility of studying of baryons as the relativistic systems of three quarks. The framework was developed for light mesons~\cite{Aniv} and baryons~\cite{Aniv,PSI}, and also for heavy-light hadrons~\cite{Manchester}. The purpose of present work is to give a description of properties of baryons containing a single heavy quark within framework proposed in ref. \cite{Aniv} and developed in ref. \cite{PSI,Manchester}. Namely, we report the calculation of observables of semileptonic decays of bottom and charm baryons: Isgur-Wise functions, asymmetry parameters, decay rates and distributions. \section{Model} Our approach \cite{Aniv} is based on the interaction Lagragians describing the transition of hadrons into constituent quarks and {\it vice versa}: \begin{eqnarray}\label{strong} {\cal L}_B^{\rm int}(x)=g_B\bar B(x)\hspace*{-0.2cm}\int \hspace*{-0.2cm} dy_1...\hspace*{-0.2cm}\int \hspace*{-0.2cm}dy_3 \delta\left(x-\frac{\sum\limits_i m_iy_i}{\sum\limits_i m_i}\right) F\left(\sum\limits_{i<j}\frac{(y_i-y_j)^2}{18}\right) J_B(y_1,y_2,y_3)+h.c.\nonumber \end{eqnarray} with $J_B(y_1,y_2,y_3)$ being the 3-quark current with quantum numbers of a baryon $B$: \begin{eqnarray}\label{current} J_B(y_1,y_2,y_3)=\Gamma_1 q^{a_1}(y_1)q^{a_2}(y_2)C\Gamma_2 q^{a_3}(y_3) \varepsilon^{a_1a_2a_3}\nonumber \end{eqnarray} Here $\Gamma_{1(2)}$ are the Dirac matrices, $C=\gamma^0\gamma^2$ is the charge conjugation matrix, and $a_i$ are the color indices. We assume that the momentum distribution of the constituents inside a baryon is modeled by an effective relativistic vertex function which depends on the sum of relative coordinates only $F\left(\frac{1}{18}\sum\limits_{i<j}(y_i-y_j)^2\right)$ in the configuration space where $y_i$ (i=1,2,3) are the spatial 4-coordinates of quarks with masses $m_i$, respectively. They are expressed through the center of mass coordinate $(x)$ and relative Jacobi coordinates $(\xi_1,\xi_2)$. The shape of vertex function is chosen to guarantee ultraviolet convergence of matrix elements. At the same time the vertex function is a phenomenological description of the long distance QCD interactions between quarks and gluons. In the case of light baryons we shall work in the limit of isospin invariance by assuming the masses of $u$ and $d$ quarks are equal each other, $m_u=m_d=m$. Breaking of the unitary SU(3) symmetry is taken into account via a difference of strange and nonstrange quark masses $m_s-m\neq 0$. In the case of heavy-light baryonic currents we suppose that heavy quark is much larger than light quark $(m_Q\gg m_{q_1},m_{q_2})$, i.e. a heavy quark is in the c.m. of heavy-light baryon. Now we discuss the model parameters. First, there are the baryon-quark coupling constants and the vertex function in the Lagrangian ${\cal L}_B^{\rm int}(x)$. The coupling constant $g_B$ is calculated from {\it the compositeness condition} that means that the renormalization constant of the baryon wave function is equal to zero, $Z_B=1-g_B^2\Sigma^\prime_B(M_B)=0$, with $\Sigma_B$ being the baryon mass operator. The vertex function is an arbitrary function except that it should make the Feynman diagrams ultraviolet finite, as we have mentioned above. We choose in this paper a Gaussian vertex function for simplicity. In Minkowski space we write $F(k^2_1+k^2_2)=\exp[(k^2_1+k^2_2)/\Lambda_B^2]$ where $\Lambda_B$ is the Gaussian range parameter which is related to the size of a baryon. It was found \cite{PSI} that for nucleons $(B=N)$ the value $\Lambda_N=1.25$ GeV gives a good description of the nucleon static characteristics (magnetic moments, charge radii) and also form factors in space-like region of $Q^2$ transfer up to 1 GeV$^2$. In this work we will use the value $\Lambda_{B_q}\equiv\Lambda_N=1.25$ GeV for light baryons and consider the value $\Lambda_{B_Q}$ for the heavy-light baryons as an adjustable parameter. As far as the quark propagators are concerned we shall use the standard form of light quark propagator with a mass $m_q$ \begin{eqnarray}\label{Slight} <0|{\rm T}(q(x)\bar q(y))|0>= \int{d^4k\over (2\pi)^4i}e^{-ik(x-y)}S_q(k), \,\,\,\,\,\, S_q(k)={1\over m_q-\not\! k}\nonumber \end{eqnarray} and the form \begin{eqnarray}\label{Sheavy} S(k+v\bar\Lambda_{\{q_1q_2\}})= \frac{(1+\not\! v)}{2(v\cdot k+\bar\Lambda_{\{q_1q_2\}}+i\epsilon)} \nonumber \end{eqnarray} for heavy quark propagator obtained in the heavy quark limit (HQL) $m_Q\to\infty$. The notation are the following: $\bar\Lambda_{\{q_1q_2\}}=M_{\{Qq_1q_2\}}-m_Q$ is the difference between masses of heavy baryon $M_{\{Qq_1q_2\}}\equiv M_{B_Q}$ and heavy quark $m_Q$ in the HQL, $v$ is the four-velocity of heavy baryon. It is seen that the value $\bar\Lambda_{\{q_1q_2\}}$ depends on a flavor of light quarks $q_1$ and $q_2$. Neglecting the SU(2)-isotopic breaking gives three independent parameters: $\bar\Lambda\equiv\bar\Lambda_{uu}=\bar\Lambda_{dd}=\bar\Lambda_{du}$, $\bar\Lambda_{s}\equiv\bar\Lambda_{us}=\bar\Lambda_{ds}$, and $\bar\Lambda_{ss}$. Of course, the deficiency of such a choice of light quark propagator is lack of confinement. This could be corrected by changing the analytic properties of the propagator. We leave that to a future study. For the time being we shall avoid the appearance of unphysical imaginary parts in the Feynman diagrams by restricting the calculations to the following condition: the baryon mass must be less than the sum of constituent quark masses $M_B<\sum\limits_i m_{q_i}$. In the case of heavy-light baryons the restriction $M_B<\sum\limits_i m_{q_i}$ trivially gives that the parameter $\bar\Lambda_{\{q_1q_2\}}$ must be less than the sum of light quark masses $\bar\Lambda_{\{q_1q_2\}} < m_{q_1}+m_{q_2}$. The last constraint serves as the upper limit for a choice of parameter $\bar\Lambda_{\{q_1q_2\}}$. Parameters $\Lambda_{B_Q}$, $m_s$, $\bar\Lambda$ are fixed in this paper from the description of data on $\Lambda^+_c\to\Lambda^0+e^+ +\nu_e$ decay. It is found that $\Lambda_Q$=2.5 GeV, $m_s$=570 MeV and $\bar\Lambda$=710 MeV. Parameters $\bar\Lambda_s$ and $\bar\Lambda_{\{ss\}}$ cannot be adjusted at this moment since the experimental data on the decays of heavy-light baryons having the strange quarks (one or two) are not available. In this paper we use $\bar\Lambda_s=$850 MeV and $\bar\Lambda_{\{ss\}}=$1000 MeV. \section{Results} In this section we give the numerical results for the observables of semileptonic decays of bottom and charm baryons: the baryonic Isgur-Wise functions, decay rates and asymmetry parameters. We check that $\xi_1$ and $\xi_2$ functions are satisfied to the model-independent Bjorken-Xu inequalities. Also the description of the $\Lambda^+_c\to\Lambda^0+e^+ +\nu_e$ decay which was recently measured by CLEO Collaboration \cite{CLEO} is given. In what follows we will use the following values for CKM matrix elements: $|V_{bc}|$=0.04, $|V_{cs}|$=0.975. In our calculations of heavy-to-heavy matrix elements we are restricted only by one variant of three-quark current for each kind of heavy-light baryon: {\it Scalar current} for $\Lambda_Q$-type baryons and {\it Vector current} for $\Omega_Q$-type baryons \cite{Shuryak,Manchester}. The functions $\zeta$ and $\xi_1$ have the upper limit $\Phi_0(\omega)=\frac{\ln(\omega+\sqrt{\omega^2-1})}{\sqrt{\omega^2-1}}$. It is easy to show that $\zeta(\omega)=\xi_1(\omega)=\Phi_0(\omega)$ when $\bar\lambda=0$. The radii of $\zeta$ and $\xi_1$ have have the lower bound $\zeta\geq 1/3$ and $\xi_1\geq 1/3$. Increasing of the $\bar\lambda$ value leads to the suppression of IW-functions in the physical kinematical region for variable $\omega$. The IW-functions $\xi_1$ and $\xi_2$ must satisfy two model-independent Bjorken-Xu inequalities \cite{Xu} derived from the Bjorken sum rule for semileptonic $\Omega_b$ decays to ground and low-lying negative-parity excited charmed baryon states in the HQL \begin{eqnarray} & &1\geq B(\omega)=\frac{2+\omega^2}{3}\xi_1^2(\omega)+ \frac{(\omega^2-1)^2}{3}\xi_2^2(\omega) +\frac{2}{3}(\omega-\omega^3)\xi_1(\omega)\xi_2(\omega) \label{ineq1}\\ & &\rho^2_{\xi_1}\geq \frac{1}{3}-\frac{2}{3}\xi_2(1) \label{ineq2} \end{eqnarray} \noindent The inequality (\ref{ineq2}) for the slope of the $\xi_1$-function is fulfilled automatically because of $\rho^2_{\xi_1} \geq 1/3$ and $\xi_2(1) > 0$. From the inequality (\ref{ineq1}) one finds the upper limit for the function $\xi_1(\omega)$: $\xi_1(\omega)\leq\sqrt{3/(2+\omega^2)}$ In Fig.1 we plot the $\zeta$ function in the kinematical region $1\leq \omega \leq \omega_{max}$. For a comparison the results of other phenomenological approaches are drawn. There are data of QCD sum rule~\cite{Grozin}, IMF models~\cite{Kroll,Koerner2}, MIT bag model~\cite{Zalewski}, a simple quark model (SQM)~\cite{Mark1} and the dipole formula~\cite{Koerner2}. Our result is close to the result of QCD sum rules~\cite{Grozin}. In Table 1 our results for total rates are compared with the predictions of other phenomenological approaches: constituent quark model \cite{DESY}, spectator quark model \cite{Singleton}, nonrelativistic quark model \cite{Cheng}. \newpage \begin{center} {\bf Table 1.} Model Results for Rates of Bottom Baryons (in $10^{10}$ sec$^{-1}$)\\ \end{center} \begin{center} \def1.{1.} \begin{tabular}{|c|c|c|c|c|} \hline Process & Ref. \cite{Singleton} & Ref. \cite{Cheng} & Ref. \cite{DESY} & Our results\\ \hline\hline $\Lambda_b^0\to\Lambda_c^+ e^-\bar{\nu}_e$ & 5.9 & 5.1 & 5.14 & 5.39 \\ \hline $\Xi_b^0\to\Xi_c^+ e^-\bar{\nu}_e$ & 7.2 & 5.3 & 5.21& 5.27 \\ \hline $\Sigma_b^+\to\Sigma_c^{++} e^-\bar{\nu}_e$ & 4.3 & & & 2.23 \\ \hline $\Sigma_b^{+}\to\Sigma_c^{\star ++} e^-\bar{\nu}_e$ & & & &4.56 \\ \hline $\Omega_b^-\to\Omega_c^0 e^-\bar{\nu}_e$ & 5.4 & 2.3 & 1.52 & 1.87\\ \hline $\Omega_b^-\to\Omega_c^{\star 0} e^-\bar{\nu}_e$ & & & 3.41 & 4.01 \\ \hline\hline \end{tabular} \end{center} \vspace*{0.4cm} Now we consider the heavy-to-light semileptonic modes. Particular the process $\Lambda^+_c\to\Lambda^0+e^++\nu_e$ which was recently investigated by CLEO Collaboration \cite{CLEO} is studied in details. At the HQL ($m_C\to\infty$), the weak hadronic current of this process is defined by two form factors $f_1$ and $f_2$ \cite{DESY,Cheng}. Supposing identical dipole forms of the form factors (as in the model of K\"{o}rner and Kr\"{a}mer \cite{DESY}), CLEO found that $R=f_2/f_1=$-0.25$\pm$0.14$\pm$0.08. Our form factors have different $q^2$ dependence. In other words, the quantity $R=f_2/f_1$ has a $q^2$ dependence in our approach. In Fig.10 we plot the results for $R$ in the kinematical region $1\leq \omega \leq \omega_{max}$ for different magnitudes of $\bar\Lambda$ parameter. Here $\omega$ is the scalar product of four velocities of $\Lambda_c^+$ and $\Lambda^0$ baryons. It is seen that growth of the $\bar\Lambda$ leads to the increasing of ratio $R$. The best fit of experimental data is achieved when our parameters are equal to $m_s=$570 MeV, $\Lambda_Q=$2.5 GeV and $\bar\Lambda=$710 MeV. In this case the $\omega$-dependence of the form factors $f_1$, $f_2$ and their ratio $R$ are drawn in Fig.11. Particularly, we get $f_1(q^2_{max})$=0.8, $f_2(q^2_{max})$=-0.18, $R$=-0.22 at zero recoil ($\omega$=1 or q$^2$=q$^2_{max}$) and $f_1(0)$=0.38, $f_2(0)$=-0.06, $R$=-0.16 at maximum recoil ($\omega=\omega_{max}$ or $q^2$=0). One has to remark that our results at $q^2_{max}$ are closed to the results of nonrelativistic quark model \cite{Cheng}: $f_1(q^2_{max})$=0.75, $f_2(q^2_{max})$=-0.17, $R$=-0.23. Also our result for $R$ weakly deviate from the experimental data \cite{CLEO} $R=-0.25 \pm 0.14 \pm 0.08$ and the result of nonrelativistic quark model (Ref. \cite{Cheng}). Our prediction for the decay rate $\Gamma(\Lambda^+_c\to\Lambda^0e^+\nu_e)$=7.22$\times$ 10$^{10}$ sec$^{-1}$ and asymmetry parameter $\alpha_{\Lambda_c}$=-0.812 also coincides with the experimental data $\Gamma_{exp}$=7.0$\pm$ 2.5 $\times$ 10$^{10}$ sec$^{-1}$ and $\alpha_{\Lambda_c}^{exp}$=-0.82$^{+0.09+0.06}_{-0.06-0.03}$ and the data of Ref. \cite{Cheng} $\Gamma$=7.1 $\times$ 10$^{10}$ sec$^{-1}.$ One has to remark that the success in the reproducing of experimental results is connected with the using of the $\Lambda^0$ three-quark current in the $SU(3)$-flavor symmetric form. By analogy, in the nonrelativistic quark model \cite{Cheng} the assuming the $SU(3)$ flavor symmetry leads to the presence of the flavor-suppression factor $N_{\Lambda_c\Lambda}=1/\sqrt{3}$ in matrix element of $\Lambda_c^+\to\Lambda^0 e^+\nu_e$ decay. If the $SU(3)$ symmetric structure of $\Lambda^0$ hyperon is not taken into account the predicted rate for $\Lambda_c^+\to\Lambda^0 e^+\nu_e$ became too large (see, discussion in ref. \cite{DESY,Cheng}). Finally, in Table 2 we give our predictions for some modes of semileptonic heavy-to-lights transitions. Also the results of other approaches are tabulated. \vspace*{0.4cm} \begin{center} {\bf Table 2.} Heavy-to-Light Decay Rates (in 10$^{10}$ s$^{-1}$). \end{center} \begin{center} \begin{tabular}{|c|c|c|c|c|c|c|} \hline Process & Quantity & Ref.\cite{Singleton} & Ref.\cite{Cheng} & Ref.\cite{Datta} & Our & Experiment \\ \hline\hline $\Lambda_c^+\to\Lambda^0 e^+\nu_e$ & $\Gamma$ & 9.8 & 7.1 & 5.36 & 7.22 & 7.0$\pm$ 2.5 \\ \hline $\Xi_c^0\to\Xi^- e^+\nu_e$ & $\Gamma$ & 8.5 & 7.4 & & 8.16 & \\ \hline $\Lambda_b^0\to p e^-\bar\nu_e$ & $\Gamma/|V_{bu}|^2$ & & & 6.48$\times$ 10$^2$ & 7.47$\times$ 10$^2$ &\\ \hline $\Lambda_c^+\to ne^+\nu_e$ & $\Gamma/|V_{cd}|^2$ & & & & 0.26$\times$ 10$^2$ & \\ \hline\hline \end{tabular} \end{center} \vspace*{.5cm} \section{Acknowledgements} We would like to thank J\"{u}rgen K\"{o}rner and Peter Kroll for useful discussions. This work was supported in part by the INTAS Grant 94-739, the Heisenberg-Landau Program by the Russian Fund of Fundamental Research (RFFR) under contract 96-02-17435-a and the State Committee of the Russian Federation for Education (project N 95-0-6.3-67, Grand Center at S.-Petersburg State University).
proofpile-arXiv_065-409
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} \no Phase ordering kinetics, critical and low temperature dynamics of pure and random systems are the subject of active research \cite{Brayrev}. Of particular interest are the approximate methods to deal with non linear dynamical equations, which often amount to a self-consistent resummation of perturbation theory \cite{BCKM}. A much debated case is the `mode-coupling' approximation, used to describe liquids approaching their frozen (glass) phase. Interestingly, this mode-coupling approximation for systems without disorder can alternatively be seen as the exact equations for an associated {\it disordered} model of the spin-glass type \cite{Krai,frhz,BCKM}. The simplest mode coupling approximation for the $\f^4$ theory is however not very good. For example, it predicts for the static critical exponent $\eta$ the value $2 - \frac{d}{2} $ independently of the number $n$ of components of the field $\f$. Furthermore, the underlying disordered model is not stable \cite{BCKM}. A better behaved resummation scheme is the ``Self-Consistent Screening Approximation" (SCSA) introduced by Bray in the context of the static $\f^4$ theory \cite{bray1,bray2}, and used in other contexts \cite{MY,LD}. It amounts to resumming self consistently all the diagrams appearing in the large-$n$ expansion, including those of order $1 \over n$. Again, this approximation becomes exact for a particular mean-field like spin-glass model \cite{BCKM}, which turns out to be well defined for all temperatures and thus ensures that the approximation is well behaved. The aim of the present paper is to generalize the SCSA equations to describe the dynamics of the $\f^4$ theory {\it at the critical point}, and to predict a value for the dynamical exponent $z$. In section \ref{scsa} we shall introduce the dynamical SCSA and the dynamical equations in their general form. From section \ref{statics} and throughout the rest of the paper we assume that time-translation invariance (TTI) and the fluctuation-dissipation theorem hold at least down to the critical point. Bray's equations will be recovered as the static limit of our dynamical equations. The reliability of the SCSA is discussed quantitatively in the $0$-dimensional static case. In section \ref{dynamics} we study the equations right at the critical temperature where dynamical scaling is supposed to hold. The full solution of these coupled equations, involving {\it scaling functions} gives in principle the dynamical exponent $z$ within the SCSA approximation. Unfortunately, as is often the case \cite{KPZ}, these equations are very hard to solve, either analytically or even numerically. In section \ref{ansatz1} and \ref{ansatz2} we thus propose two different ans\"atze for the scaling functions, which are however much constrained by general requirements. The second ansatz leads to the exact $O(\eps^2)$ result in the $\eps=4-d$ RG expansion of Halperin, Hohenberg and Ma \cite{HHM}. The numerical value of the exponent $z$ only very weakly depends on the chosen ansatz, and turns out to be quite close to the best available Monte-Carlo estimate for the Ising model in $d=3$ \cite{Heuer}. \section{The Self Consistent Screening Approximation} \label{scsa} Let us consider the coarse-grained Hamiltonian density \be {\cal H}[{\bf \f} (\vec{x})] = \frac{1}{2}(\nabla \f (\vec{x}))^2 + \frac{\mu}{2} \f^2 (\vec{x}) - \frac{g}{8} \f^4 (\vec{x}), \label{h1} \ee \no where ${ \f(\vec{x}) }$ is an $n$ component field and $\vec{x}$ is the $d$ dimensional space variable. With $\f^2 (\vec{x})$ and $\f^4 (\vec{x})$ we indicate respectively $|{\f(\vec{x}) }|^2$ and $(|{ \f(\vec{x}) }|^2 )^2$. The coupling constant $g$ is negative and of order $n^{-1}$; $\mu$ is a (temperature dependent) mass term which vanishes at the mean-field transition point. The partition function is \be Z = \int {\cal D} {\bf \f} e^{-\int d^{d}x \ \frac{{\cal H}[{\bf \f} (\vec{x})]}{T}}, \ee \no In order to introduce the Self Consistent Screening Approximation one starts from a large-$n$ expansion formalism. We re-write $Z$ with a gaussian transformation introducing an auxiliary field $\s$ \be Z = \int {\cal D} \s {\cal D} {\bf \f} e^{-\int d^{d}x \ \frac{ {\cal H}[{\bf \f} (\vec{x}),\s(\vec{x})]}{T}} , \ee \no $H[{\bf \f} (\vec{x}),\s(\vec{x})]$ being now the Hamiltonian density of two coupled fields ${\bf \f} (\vec{x})$ and $\s(\vec{x})$. \be H[\s,{\bf \f} ] = \frac{1}{2}(\nabla \f (\vec{x}))^2 + \frac{\mu}{2} \f^2 (\vec{x}) + \frac{1}{2}\s^2 (\vec{x}) - \frac{\sqrt{g}}{2} \s (\vec{x}) \f^2 (\vec{x}) . \ee The SCSA amounts to consider the renormalization of the order $1/n$ diagrams in the Dyson expansion for the correlation functions of the two fields ${\bf \f} (\vec{x})$ and $\s(\vec{x})$. Using this resummation scheme Bray \cite{bray1} obtained interesting results for the static exponent $\n$ which describes the small momentum behaviour of the correlation functions. The static SCSA equations for $\langle {\bf \f} (\vec{x}) {\bf \f} (\vec{x}^{\prime}) \rangle$ (plain line) and $\langle \s(\vec{x}) \s(\vec{x}^{\prime}) \rangle$ (``dashed" line) are reported diagrammatically in figure 1. The bare quantities are indicated respectively with a thinner plain line and with a dashed line. \begin{figure} \epsfbox{diagst.ps} \vspace{0.5cm} \caption[]{} \end{figure} Our goal is to develop a dynamical generalization of this expansion for non-conserved Langevin dynamics, starting from the SCSA Hamiltonian. We thus obtain the following equations of motion for ${\bf \f} (\vec{x},t)$ and $\s(\vec{x},t)$: \begin{eqnarray} \dot{{\bf \f} }(\vec{x},t)& = &-(\nabla^2 + \mu) {\bf \f} (\vec{x},t) + \sqrt{g} {\bf \f} (\vec{x},t) \s (\vec{x},t) + \eta_{\f} (\vec{x},t) \label{langf} \\ \dot{\s}(\vec{x},t)& = & - \s (\vec{x},t) + \frac{\sqrt{g}}{2} \f^2(\vec{x},t) + \eta_{\s}(\vec{x},t). \label{langs} \end{eqnarray} \noindent with two independent thermal noises $\eta_{\f}, \eta_{\s}$. Let us now consider the two-point functions \begin{eqnarray} G_{\f}(\vec{x},\vec{x}^{\prime},t,t^{\prime}) &=& \left< \frac{\partial{{\bf \f}(\vec{x},t)}} {\partial{\n(\vec{x}^{\prime},t ^{\prime})}} \right> \\ C_{\f}(\vec{x},\vec{x}^{\prime},t,t^{\prime}) &=& < {\bf \f} (\vec{x},t) {\bf \f}(\vec{x}^{\prime},t^{\prime}) >, \end{eqnarray} and the corresponding functions for the field $\s$. The SCSA dynamical equations, which can be seen as a Mode-Coupling approximation on the set of equations (2.5-6) (see figure 2) then read: \begin{eqnarray} \nonumber \Sigma_{\f}(t_1,t_2) & = & n \frac{g}{2} \delta(t_1-t_2) \int_{0}^{t_{1}} dt_{3} C_{\f}(t_{3} ,t_{3}) G_{\s}^{0} (t_{1},t_{3}) + \\ &+& g [G_{\f}(t_{1},t_{2})C_{\sigma}(t_{1},t_{2}) + G_{\s}(t_{1},t_{2})C_{\f}(t_{1},t_{2})] \label{2times1} \end{eqnarray} \be \Sigma_{\s}(t_1,t_2) = n g G_{\f}(t_{1},t_{2}) C_{\f}(t_{1},t_{2}) \ee \begin{eqnarray} D_{\f}(t_1,t_2) &=& 2T \delta (t_1-t_2) + g C_{ \f}(t_1,t_2) C_{ \s}(t_1,t_2) \\ D_{\s}(t_1,t_2) &=& 2T \delta (t_1-t_2) + n \frac{g}{2} C_{\f}^2 (t_1,t_2), \label{2times} \end{eqnarray} where we have dropped the space coordinates $\vec x$ for clarity, and introduced the self-energies $\Sigma$, defined as: \be G(t,t^{\prime}) = G^0(t,t^{\prime}) + \int_{0}^{t} dt_{1} \int_{t'}^{t_1} dt_2 G^{0}(t,t_{1}) \Sigma(t_{1},t_{2}) G (t_{2},t^{\prime}), \ee (the label $^0$ refers to the bare quantity), and the `renormalized noises' $D$, defined as: \be C(t,t^{\prime}) = \int_{0}^{t}dt_{1} \int_{0}^{t^{\prime}}dt_{2} G(t,t_{1})D(t_{1},t_{2})G(t^{\prime},t_{2}). \label{cdef} \ee \begin{figure} \epsfbox{diagdy.ps} \vspace{0cm} \caption[]{} \end{figure} We shall limit ourselves to consider the above equations in a regime of stationary dynamics. That is to say that we will make use of the assumption of time translational invariance (only differences of times matter), which allows one to show that the fluctuation dissipation theorem (FDT) is valid, i.e: \be \theta(t-t^{\prime}) \frac{\partial C(t-t^{\prime})}{\partial{t^{\prime}}} = T G(t-t^{\prime}). \ee Extensions of these methods to non stationary low temperature regime, where this theorem is violated \cite{CK}, will be subject of further work. In the following, we shall set the energy scales by choosing $T=1$, and vary the mass term $\mu$ to reach the critical point. \section{Static Limit} \label{statics} With these assumptions equations (\ref{2times}) reduce to only two coupled independent equations which have the simplest form in Fourier space \begin{eqnarray} \nonumber \Sigma_{\f}(k,\w) &=& g \int \left[ C_{\s}(k- k^{\prime}, \w - \w^{\prime}) G_{\f}(k^{\prime},\w^{\prime}) + C_{\f}(k-k^{\prime},\w - \w^{\prime}) G_{\s}(k^{\prime},\w^{\prime}) \right] Dk'D\w' \\ && + \frac{ n g}{2} G_{\s}^{0}(k=0,\w=0) \int C_{\f}(k^{\prime},\w^{\prime}) Dk'D\w' \label{sigf} \\ \Sigma_{\s}(k,\w) &=& {n g} \int C_{\f}(k- k^{\prime},\w - \w^{\prime}) G_{\f}(k^{\prime},\w^{\prime}) Dk'D\w' . \label{sigs} \end{eqnarray} where $Dk' \equiv \frac{d^d k'}{(2\pi)^d}$ and $D\w' \equiv \frac{d\w'}{2\pi}$. Using the fact that $C(k,t=0) \equiv {\cal C}(k)$ is equal to $G(k,\omega=0)$ (from FDT and the Kramers-Kronig (KK) relations), and using again the KK relations, it is easy to check that for $\w=0$ one recovers exactly the static SCSA equations \cite{bray1}, namely \begin{eqnarray} {\cal C}_{\f}(k) & = & \frac{1}{\mu + k^2 - g \int Dk' {\cal C}_{\f}(k-k'){\cal C}_{\s}(k') -\frac{g n}{2} \int Dk' {\cal C}_{\f}(k')} \nonumber \\ {\cal C}_{\s}(k) & = & \frac{1}{1 - \frac{g}{2} n \int Dk'{\cal C}_{\f}(k-k'){\cal C}_{\f}(k')}, \label{cstatiche} \end{eqnarray} \noindent In order to test the validity of this approximation, it is interesting to consider the case of zero spatial dimensions \cite{bray2}. Let us set $n=1$ which is a bad case for the SCSA which should become more accurate the larger $n$ is. We will compare equations (\ref{cstatiche}) with the exact static correlation function which in zero dimension can be calculated analytically and is \be {\cal C}_{exact} = -{1\over { \mu}} + {\mu \over g} - {{\mu \,{\rm K}_{-\frac{3}{4}}( {{{\mu^2} }\over {4\,g}})}\over {2\,g\,{\rm K}_{1\over 4}( {{- \,{\mu^2} }\over {4\,g}})}} - {{\mu \,{\rm K}_{5\over 4}( {{{\mu^2} }\over {4\,g}})}\over {2\,g\,{\rm K}_{1\over 4}( {{- \,{\mu^2} }\over {4\,g}})}}, \ee \no where $K_{n}(a)$ is the modified Bessel function of the second kind. Equations (\ref{cstatiche}) give for ${\cal C}_{\f}$: \be {\cal C}_{scsa} = \frac{1}{\left(\mu - n \frac{g}{2} {\cal C}_{scsa} - g \frac{{\cal C}_{scsa}} { \left( 1 - \frac{g}{2} n {\cal C}_{scsa}^2 \right)}\right)} . \ee From plotting the relative difference of the two correlation functions versus the coupling (see figure 3) we can see that SCSA is quite close to the exact theory. In particular, the asymptotic behaviour in the $|g| \rightarrow \infty$ limit of the two functions is \be \lim_{|g| \rightarrow \infty} \sqrt{|g|} {\cal C}_{scsa} = 2(\sqrt{2} - 1) {\mbox{ and }} \lim_{|g| \rightarrow \infty} \sqrt{|g|} C_{exact} = {{2\,{\sqrt{2}}\,{\Gamma}({3\over 4})}\over {{\Gamma}({1\over 4})}}. \ee For all $g$, the relative difference is actually bounded by: \be \frac{\left| {\cal C}_{exact} - {\cal C}_{scsa} \right|} {{\cal C}_{exact}} < 1 - {{{\sqrt{-1 + {\sqrt{2}}}}\,{\rm \Gamma}({1\over 4})}\over {2\,{\rm \Gamma}({3\over 4})}} = 0.0479... \ee \no We can also compare the small$-g$ expansions of the two theories which give \begin{eqnarray} {\cal C}_{exact} &=& \frac{1}{ \mu}(1 + \frac{3}{2 \mu^2} g + \frac{21}{4 \mu^4} g^2) \\ {\cal C}_{scsa} &=& \frac{1}{ \mu}(1 + \frac{3}{2 \mu^2} g + \frac{5} {\mu^4} g^2) \end{eqnarray} showing explicitly how the two theories differ already at order $g^{2}$. The self consistent nature of the approximation however keeps the SCSA in good agreement with the exact theory even for large values of the coupling constant as remarked before. It is instructive, in passing, to compare the SCSA with the simple Hartree ($n = \infty$) resummation scheme, which is also the Gaussian variational result. One defines $F_{H} = \min{\{F\} }$ where \be F = F_{0} + <H-H_{0}>, \ee with \be F_{0} = - \ln \int {\cal D} {\bf \f} e^{-\frac{\tilde{\mu}\f^2}{2}} = - \ln{(\frac{2 \pi}{\tilde{\mu}})} \ee \be <H_{0}> = \frac{1}{2 } \ee \be <H> = \int {\cal D} {\bf \f} e^{-\frac{\tilde{\mu}\f^2}{2}} \left( \frac{\mu}{2}\f^{2} - \frac{g}{8}\f^{4} \right) = \left( \frac{\mu}{2\tilde{\mu}} - \frac{3g}{8 \tilde{\mu}^{2}} \right) \ee Minimising $F$ with respect to $\tilde{\mu}$ we find \be \mu_{H} = \frac{\mu + \sqrt{\mu^2 - 6g}}{2}, \ee and consequently \be {\cal C}_{H} = <\f^2>_{\mu_{H}} = \frac{2}{\mu + \sqrt{\mu^2 - 6g}}. \ee As can be seen from figure 3, the SCSA turns out to be fairly better than the Hartree variational approach (at least in this particular case of $n=1$ and $d=0$). \begin{figure} \epsfbox{comparison.ps} \vspace{0.5cm} \caption[]{Relative difference between the exact result and the Hartree (${\cal C}_H$) and the SCSA (${\cal C}_{scsa}$) approximations, in the case $n=1,d=0$.} \end{figure} \section{Critical Dynamics} \label{dynamics} We shall now work right at the critical point $\mu_{c}$ such that the renormalised mass vanishes (therefore eliminating the `tadpole' contribution in Eq. \ref{2times1}). We shall search for solutions under the general dynamic scaling form (valid in the small-$k$ and small-$\w$ limit): \begin{eqnarray} G_{\f}(k,\w) &=& \frac{1}{k^{\Delta}} n_{\f}(\frac{\w}{k^{z}}) \hspace{2truecm} G_{\s}(k,\w) = \frac{1}{k^{\Delta^{\prime}}} n_{\s}(\frac{\w}{k^{z}}) \nonumber \\ C_{\f}(k,\w) &=& \frac{2}{ \w k^{\Delta}} Im \left[ n_{\f}(\frac{\w}{k^{z}}) \right ] \hspace{.7truecm} C_{\s}(k,\w) = \frac{2}{ \w k^{\Delta^{\prime}}} Im \left[ n_{\s}(\frac{\w}{k^{z}}) \right ]. \label{generalscaling} \end{eqnarray} where we have defined $\Delta = 2- \n$, and used FDT. \no Setting first $\w=0$, one finds by matching the momentum dependence of the left and right hand sides of (\ref{sigf}-\ref{sigs}) that: \be \Delta^{\prime} = d - 2 \Delta = d - 4 + 2 \n. \ee Note that in mean field, $z=2$, $\Delta =2$, $\eta=0$ and $\Delta'=0$. Identification of the prefactors yields: \be n_{\s}(0) n_{\f}^2 (0) = - \frac{2}{ f(\eta,d) n g } \label{relstat} \ee where \be f(\eta,d) = \frac{1}{(4 \pi)^{d/2}} \frac{\Gamma[\Delta - \frac{d}{2}] \Gamma[\frac{d-\Delta}{2}]^2 }{\Gamma[d-\Delta] \Gamma[\frac{\Delta}{2}]^2} . \label{fbray} \ee and an extra equation fixing $\eta$ as a function of $d$ and $n$, which we do not write explicitely \cite{bray1}. Now let us consider the other case where $k=0$ and $\w > 0$ (but small). Taking the imaginary part of (\ref{sigf}-\ref{sigs}), one obtains: \begin{eqnarray} Im \left[ \Sigma_{\f}(0,\w) \right] &=& \frac{S \w }{n n_{\f} (0)} \int q^{\Delta -1} dq ds \frac{Im\left[ f_{\f} (\frac{(\w -s)}{q^{z}}) \right] Im\left[ f_{\s} (\frac{s}{q^{z}}) \right] }{s (\w -s)} \label{fk=0} \\ Im \left[ \Sigma_{\s}(0,\w)\right] &=& \frac{S}{n_{\s}(0)} \int q^{\Delta ^{\prime} -1} dq d s \frac{Im\left[ f_{\f} (\frac{(\w -s)}{q^{z}}) \right] Im\left[ f_{\f} (\frac{(s)}{q^{z}}) \right] }{s}, \label{sk=0} \end{eqnarray} \no where $f_{\f,\s}(x) = n_{\f,\s}(x)/n_{\f,\s}(0)$. We also defined \be S = \frac{2 n g \Omega_{d}}{(2 \pi)^{(d+1)}} n_{\f}^2(0) n_{\s}(0) \equiv -\frac{4 \Omega_{d}}{f(\eta,d)(2 \pi)^{(d+1)}} \label{eq1} \ee In general the scaling functions can be written \begin{eqnarray} Im[f_{\f}(x)] & \doteq & A \tilde{f_{\f}} (a x) \\ \nonumber Im[f_{\s}(x)] & \doteq & A'\tilde{f_{\s}} (a^{\prime} x), \end{eqnarray} \no with by convention $\lim_{u \to \infty}u^{\Delta/z} \tilde{f_{\f}}(u) =1$ and $\lim_{u \to \infty} u^{\Delta'/z} \tilde{f_{\s}}(u)=1$. This asymptotic behaviour is required for the $k \to 0$ limit to be well defined, if (\ref{generalscaling}) is correct. Furthermore, the small-$\w$ behaviour of the imaginary part of the response function is expected to be regular for $k$ finite, and hence $\tilde{f}(u) \propto u$ for $u \to 0$. $A,A'$ are coefficients setting the scale of the imaginary part of the response function while $a,a'$ are coefficients setting the frequency scales. Using the fact that the imaginary and real part of the response function are power-laws at large frequencies, which imply that their ratio is $\tan\left( \frac{\pi \Delta}{2z} \right)$ (resp. $\tan\left( \frac{\pi \Delta'}{2z} \right)$), one finds that: \begin{eqnarray} \nonumber \frac{a^{\Delta/z} }{A} \sin^{2} \left( \frac{\pi \Delta}{2z} \right) &=& \frac{ S }{nz} \int_{0}^{\infty} \frac{dx}{x^{1 + \Delta/z}} \int_{-\infty}^{\infty} \frac{du}{u(1-u)} Im\left[ f_{\f} (x (1-u)) \right] Im\left[ f_{\s}(x u) \right] \\ \frac{\alpha'^{\Delta^{\prime}/z} }{A'} \sin^{2} \left( \frac{\pi \Delta^{\prime}}{2z} \right) &=& \frac{ S }{z} \int_{0}^{\infty} \frac{dx}{x^{1 + \Delta^{\prime}/z}} \int_{-\infty}^{\infty} \frac{du}{u} Im\left[ f_{\f} (x (1-u)) \right] Im\left[ f_{\f} (x u) \right] \label{scsagens} \end{eqnarray} It is easy to show that these equations actually only depend on the value of the {\it ratio} of frequency scales $y=\frac{a'}{a}$. The coefficient $A$ can be fixed using the KK relation, since the involved integral converges, which means that the small-$k$ behaviour of the real part of the correlation function is fully determined by the imaginary part in the scaling region $\omega, k \to 0$. Hence: \be 1 = \frac{A}{\pi} \int_{-\infty}^{\infty} dx \frac{\tilde{f_{\f}} (x )}{x}. \label{kk} \ee The corresponding integral for $\tilde{f_{\s}}$ does not converge for large $x$, meaning that the non-scaling region is needed to saturate the sum-rule. Hence, we must use another relation to fix $A'$, which we choose to be the small-$\w$ expansion of Eq. (\ref{sk=0}). Thus, {\it if} the functions $\tilde{f_{\f}},\tilde{f_{\s}}$ were known, we would have four equations to fix four constants: $A,A',y$, and, of course, the dynamical exponent $z$, in terms of $d$ and $n$. $\tilde{f_{\f}},\tilde{f_{\s}}$ are in principle also fixed by the full equations for arbitrary $\frac{\omega}{k^z}$. However, as in other similar cases \cite{KPZ}, these equations are very hard to solve, either analytically or numerically. We will thus propose ans\"atze for these functions, which have to satisfy the above general requirements. Note that once $A,A',a,a'$ have been pulled out, the only freedom is in the {\it shape} of these functions. We shall thus work with two such ans\"atze, which will turn out to give very similar answers for $z$. This was also the case in the context of the KPZ equation \cite{KPZ}. \section{Ansatz 1} \label{ansatz1} The simplest ansatz one can think of, which generalizes the mean field shape: \be \tilde{f_{\f}}(x) = \frac{x}{(1+x^2)} \label{fmf} \ee reads: \begin{eqnarray} \tilde{f_{\f}}(x) &=& \frac{x}{\left(1+x^2 \right)^{\alpha}} \\ \tilde{f_{\s}}(x) &=& \frac{x}{\left(1+x^2\right)^{\ap}} , \end{eqnarray} where we have set \begin{eqnarray} \alpha &\doteq& \frac{\Delta + z}{2z} \\ \ap &\doteq& \frac{\Delta^{\prime} + z}{2z}. \label{alphas} \end{eqnarray} (Note that $\alpha=1$ in mean field). These functions have indeed the correct asymptotic behaviours; they go linearly to zero for small values of the argument and behave as power laws ($\tilde{f_{\f}}(x) \simeq x^{-\frac{\Delta}{z}} $ and $ \tilde{f_{\s}}(x) \simeq x^{-\frac{\Delta^{\prime}}{z}}$) in the large-$x$ limit. We can now use (\ref{kk}) to determine $A$ \be A = \sqrt{\pi} \frac{\Gamma[\alpha]}{\Gamma\left[\alpha - \frac{1}{2} \right]}. \ee The small-$\w$ expansion of $Im \Sigma_{\s}(k,\w) $ can be matched with that of the right hand side of Eq.(\ref{sk=0}) leading to the following equation \be y = -\frac{2 A^2}{A' f(\eta,d) (2 \pi)^{d+1}} \int_{-\infty}^{\infty} d^d q \frac{1}{|q|^{\Delta} |1-q|^{\Delta + z}} \int_{-\infty}^{\infty} dt \frac{1}{(1+t^2)^{\alpha} \left(1+ (\frac{|q|^{z} t}{|1-q|^{z}} )^2 \right)^{\alpha}} \label{ggg} \ee After some algebraic manipulations we obtain for the last three equations: \begin{eqnarray} \nonumber \sin^{2} \left( \frac{\pi \Delta}{2z} \right) &=&- \frac{A^2 A'y S}{2 nz} B\[[1 -\frac{\Delta}{2z}, \frac{d}{2z}\]] \\& & \int_{-\infty}^{\infty} \frac{du}{ |u|^{2-\frac{\Delta}{z}}} F\[[\ap,1 -\frac{\Delta}{2z},\alpha + \ap,1- y^2 \frac{(1-u)^2}{u^2}\]] \label{scsa11} \end{eqnarray} \begin{eqnarray} \nonumber \sin^{2} \left( \frac{\pi \Delta^{\prime}}{2z} \right) &=&- \frac{A^2 A^{\prime}S}{2z} y^{-\frac{\Delta^{\prime}}{z}} B \[[1 -\frac{\Delta^{\prime}}{2z}, \frac{d}{2z}\]]\\ & & \int_{-\infty}^{\infty} \frac{u du}{|u|^{2 -\frac{\Delta^{\prime}}{z}}} F\[[\alpha,1 -\frac{\Delta^{\prime}}{2z},2 \alpha,\frac{2u-1}{u^2}\]] \label{scsa12} \end{eqnarray} \begin{eqnarray} \nonumber y &=& \pi \frac{A^2 S}{z A' \Omega_{d}} B\[[\frac{1}{2},2 \alpha -\frac{1}{2}\]] \\ & & \int_{0}^{\infty} dq q^{d-2-\Delta} \int_{|1-q|^{2z}}^{|1+q|^{2z}} \frac{dx}{x^{\frac{\Delta +3z- 2}{2z}}} F\[[\alpha,\frac{1}{2},2\alpha,1-\frac{q^{2z}}{x}\]], \label{scsa13} \end{eqnarray} where $B[a,b]$ and $F[a,b,c,x]$ are the Euler Beta and Hypergeometric functions and where the last equation (\ref{scsa13}) was written for the special case $d=3$ which we shall consider below. We can solve analytically Eqs. (\ref{ggg},\ref{scsa11},\ref{scsa12}) at order $\eps^2$ to compare with the exact RG treatment of \cite{HHM}. At lowest order we obtain: \begin{eqnarray} c &=& \frac{8 \ln{2}}{ \pi} \frac{ \arctan \sqrt{\frac{1 - y^2} {y^2}}}{\sqrt{1 - y^2}}-1 \\ A' &=& - \frac{\pi \eps}{4} \\ y &=& \frac{4 \ln 2 }{\pi} \label{oeps2} \end{eqnarray} where we have defined, following \cite{HHM}, \be z = 2 + c \eta . \label{formz} \ee The order $O(\eps^2)$ RG result reads, $c =6\ln{\frac{4}{3}} -1 =0.7261$. The form (\ref{formz}) means that to lowest order $z$ depends on $n$ only through the static exponent $\n$. On the other hand, Eqs. (\ref{oeps2}) give \be c = 0.8376, \ee in slight disagreement with the exact result. This comes from the fact that while our ansatz for $\tilde f_{\f}$ is exact in the limit $\epsilon \to 0$, the corresponding ansatz for $\tilde f_{\s}$ is already wrong at lowest order since it does not satisfy Eq. (\ref{sk=0}). In our second ansatz, we thus keep the same shape for $\tilde f_{\f}$, but choose for $\tilde f_{\s}$ a form which is exact when $\epsilon \to 0$. \section{Ansatz 2} \label{ansatz2} Knowing the mean field form for $f_{\f}(x)$ we can, at lowest order in $\epsilon$, write for $Im[f_{\s}(x)]$ \be Im f_{\s}(x) = 2^{d-4} \frac{ f(\eta,d)}{\pi^{d/2}} \frac{(2\pi)^d}{\Gamma[2-\frac{d}{2}]} Im \[[ \frac{1}{\xi(x)} \]] \label{gsgen} \ee where \be \xi(x) = 1 - \frac{\eps}{2} \int_{0}^{1} dt \log\[[ 1-t^2 - 2 i x (1-t) \]]. \ee It is then straightforward to generalize $Im[f_{\s}(x)]$ to general dimensions as: \be \tilde {f_{\s}} \propto Im \[[ (2 - i x )^{1 - \frac{\Delta^{\prime}}{z}} - (1 - i x )^{1 - \frac{\Delta^{\prime}}{z}} \]] \label{fsfin} \ee with a prefactor ensuring that the coefficient of $x^{-\frac{\Delta^{\prime}}{z}}$ for large $x$ is unity. Eq. (5.8) is now replaced by: \be \sin^2 \((\frac{\pi \Delta}{2 z} \)) = \frac{A^2 A' S b}{n z} \int_{0}^{\infty} \frac{{\mbox d} r}{r^{\frac{\Delta}{z}}} \int_{-\infty}^{\infty} {\mbox d} u \frac{Im \[[ (2 - i \frac{\pi}{2 \ln{2}} (y r u) )^{1 - \frac{\Delta^{\prime}}{z}} - (1 - i \frac{\pi}{2 \ln{2}} (y r u) )^{1 - \frac{\Delta^{\prime}}{z}} \]]}{u \[[1 + r^2 (1-u)^2 \]]^{\alpha}}. \label{scsa21} \ee where now $b$ is given by: \be b = \frac{2 \ln{2}}{\pi \((2^{ - \frac{\Delta^{\prime}}{z}} -1\)) \((\frac{\Delta^{\prime}}{z} -1 \)) }. \ee We finally obtain a set of equations for $z$ of the same kind as (\ref{scsa11}-\ref{scsa13}) but which now exact up to $O(\epsilon^2)$, as we have checked directly. \section{Numerical Results} We solved numerically both sets of equations in $d= 3$ for $n=1,...,10$. We used the values of $\eta(d=3,n)$ that can be derived from the formula reported in \cite{bray1}. The values obtained for $z$ are reported in the following table. \medskip \begin{equation} \begin{array}{|c|c|c|} \hline \par n & z \mbox{ (ansatz 1 )} & z \mbox{ (ansatz 2 )}\\ \hline 1& 2.119 & 2.113 \\ \hline 2 & 2.071 & 2.069 \\ \hline 3 & 2.050& 2.049 \\ \hline 4 & 2.038 & 2.038 \\ \hline 5 & 2.031& 2.031 \\ \hline 6 & 2.0258& 2.0258 \\ \hline 7 & 2.0223& 2.0222 \\ \hline 8 & 2.0196& 2.0195 \\ \hline 9 & 2.0174& 2.0174 \\ \hline 10 & 2.0157& 2.0157 \\ \hline \end{array} \nonumber \end{equation} \medskip As it was hoped, the results are fairly independent from the ansatz used, which is more and more true for large $n$. The result for $n=1$ is rather close to the best Monte-Carlo estimate of ref. \cite{Heuer}, which gives $z=2.09 \pm 0.02$. Let us note however that the SCSA overestimates significantly $\eta$ in $d=3$. In figure (4) we compare the two different choices for the scaling function $f_{\s}(x)$ with their relative values of the parameters $y$, $A'$ and $z$, and in the case $n=1,d=3$. We notice that the constraints for small $x$ and large $x$ restrict very much the freedom on the shape of this function. \begin{figure} \epsfbox{diff.ps} \vspace{0.5cm} \caption[]{The two ans\"atze for the functions $f_{\s}(x)$, $n=1$} \end{figure} Finally, a linear regression of our results for $n=1-10$ gives $z \simeq 2 + c \eta$ with $c=0.64$, which is lower that the $O(\eps^2)$ result, but larger that the exact result for $d=3$, $n \to \infty$, i.e. $c=\frac{1}{2}$ \cite{HHM}. \section{Conclusions} The aim of this paper was extend the static Self-Consistent screening approximation to dynamics, in particular to calculate the properties of the critical dynamics of the $\phi^4$ model. Although the resulting equations cannot be fully solved, a much constrained ansatz leads to a value of the exponent $z$ in rather good agreement with Monte-Carlo data. Our work was originally inspired by glassy dynamics: the SCSA equations actually describe in exactly the dynamics of some mean-field spin glass like models. It would be interesting to study these equations in the low temperature phase, where dynamics becomes non stationary (aging) and FDT is lost. For $\phi^4$ models, this corresponds to a coarsening regime \cite{Brayrev}. It would be interesting to know whether the SCSA equations describe properly this regime, and can compete with other approximation schemes \cite{Brayrev,NB}. \vskip 1cm {\it Acknowledgments} It is a pleasure for us to thank A.~Barrat, A.~Bray, L.F.~Cugliandolo, J.~Kurchan, E.~Maglione, M.~M\'ezard, R.~Monasson, G.~Parisi and P.~Ranieri for very instructive discussions. \bibliographystyle{IEEE}
proofpile-arXiv_065-410
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} Recent remarkable developments in superstring theory lead to the discovery that five known superstring theories in ten dimensions are related by duality transformations and that there are also $M$-theory in $11d$ and $F$-theory in $12d$ that are useful in the study of the moduli space of quantum vacua \cite {FILQ} -\cite{Vaf}. Duality requires the presence of extremal black holes in the superstring spectra. A derivation of the Bekenstein-Hawking formula for the entropy of certain extremal black holes was given using the D-brane approach \cite{SV}-\cite{pol2}. In all these developments the study of p-brane solutions of the supergravity equations play an important role \cite{DGHR}-\cite{berg}. In this paper we shall consider the following action \begin{equation} I=\frac{1}{2\kappa ^{2}}\int d^{D}x\sqrt{-g} (R-\frac{1}{2}(\nabla \phi) ^{2} -\frac{1}{2(q+1)!}e^{-\alpha \phi}F^{2}_{q+1} -\frac{1}{2(d+1)!}e^{\beta \phi}G^{2}_{d+1}), \label{011} \end{equation} It describes the interaction of the gravitation field $g_{MN}$ with the dilaton $\phi$ and with two antisymmetric fields: $F_{q+1} $ is a closed $q+1$-differential form and $G_{d+1}$ is a closed $d+1$-differential form. Various supergravity theories contain the terms from (\ref{011}). The aim of this paper is to present a solution of (\ref{011}) with the metric of the form \begin{equation} ds^{2}=H_{1}^{-2a_{1}}H_{2}^{-2a_{2}}\eta_{\mu \nu} dy^{\mu} dy^{\nu}+ H_{1}^{-2b_{1}}H_{2}^{-2a_{2}}\delta_{nm}dz^{n}dz^{m}+ H_{1}^{-2b_{1}}H_{2}^{-2b_{2}}\delta_{\alpha\beta}dx^{\alpha}dx^{\beta}, \label{12} \end{equation} where $\eta_{\mu\nu}$ is a flat Minkowski metric, $$\mu, ~\nu = 0,...,q-1;~~m,n=1,2,...,d-q,$$ and $$\alpha,\beta =1,...,D-d.$$ For definitness we assume that $ D>d\geq q$. The parameters $a_{i}$ and $b_{i}$ in the solution (\ref{12}) are rational functions of the parameters in the action (\ref{011}): \begin{equation} a_{1}=\frac{2\tilde q}{\alpha^{2}(D-2)+2q\tilde q}, ~~~~a_{2}=\frac{\alpha^{2}(D-2)} {\alpha^{2}d(D-2)+2\tilde d q^{2}} \label{13} \end{equation} \begin{equation} b_{1}=-\frac{2q}{\alpha^{2}(D-2)+2q\tilde q}, ~~~~b_{2}=-\frac{\alpha^{2}d(D-2)}{\tilde d[ \alpha^{2}d(D-2)+2\tilde d q^{2}]} \label{14} \end{equation} where \begin{equation} \tilde d=D-d-2,~~~~\tilde q=D-q-2. \label{14a} \end{equation} Our solution (\ref{12}) is valid only if the following relation between parameters in the action is satisfied \begin{equation} \alpha\beta=\frac{2q\tilde d}{D-2} \label{15} \end{equation} There are two arbitrary harmonic functions $H_{1}$ and $H_{2}$ of variables $x^{\alpha}$ in (\ref{12}), \begin{equation} \Delta H_{1}=0,~~~~~\Delta H_{2}=0. \label{16} \end{equation} Non-vanishing components of the differential form are given by \begin{equation} {\cal A}_{\mu_{1}...\mu_{q}}=h {\epsilon_{\mu_{1}...\mu_{q}}}H_{1}^{-1},~~~~ F=d{\cal A}, \label{18} \end{equation} \begin{equation} {\cal B}_{I_{1}...I_{d}}=k \epsilon_{I_{1}...I_{d}}H_{2}^{-1}, ~~~G=d{\cal B},~~~I=0,...d-1. \label{18a} \end{equation} Here $\epsilon _{123..,q}=1$, $\epsilon _{123...d}=1$ and $h$ and $k$ are given by the formulae \begin{equation} h^{2}=\frac{4(D-2)} {\alpha ^{2}(D-2)+2q\tilde q}, \label{194} \end{equation} \begin{equation} k^{2}=\frac{2\alpha^{2} (D-2)^{2}} {{\tilde d}[\alpha ^{2}d(D-2)+2q^{2}{\tilde d}]}, \label{195n} \end{equation} The dilaton field is \begin{equation} \phi=\frac{1}{2}\beta k^{2}\ln H_{2}-\frac{1}{2} \alpha h^{2}\ln H_{1}. \label{17} \end{equation} We obtain the solution (\ref{12}) by reducing the Einstein equations to the system of algebraic equations. To this end we introduce a linear dependence between functions in the Ansatz (see below). The solution (\ref{12}) consists of three blocks, the first block consists of variables $y$, another of variables $z$ and the other of variables $x$ and all functions depend only on $x$. We shall call it the three-block p-brane solution. We shall consider also the following "dual" action \begin{equation} \tilde I=\frac{1}{2\kappa ^{2}}\int d^{D}x\sqrt{-g} (R-\frac{1}{2}(\nabla \phi) ^{2} -\frac{1}{2(q+1)!}e^{-\alpha \phi}F^{2}_{q+1} -\frac{1}{2({\tilde d}+1)!}e^{\tilde\beta \phi}G^{2}_{{\tilde d}+1}), \label{11} \end{equation} where $G_{\tilde d+1}$ is a closed $\tilde d+1$-differential form. If $\tilde{d}$ is related to $d$ by (\ref {14a}) and \begin{equation} \tilde\beta =-\beta \label{11a} \end{equation} then the solution for the metric (\ref{12}) with the differential form $F$ (\ref{18}) and the dilaton (\ref{17}) is valid also for the action (\ref{11}). An expression for the antisymmetric field $G$ will be different, namely \begin{equation} G^{\alpha_{1}...\alpha _{\tilde d+1}}=k H_{1}^{\sigma_{1}}H_{2}^{\sigma_{2}} \epsilon ^{\alpha _{1}...\alpha_{\tilde d+1}\beta} \partial _{\beta} H_{2}^{-1}. \label{1999} \end{equation} here $\epsilon ^{123..\tilde d+2}=1$ and \begin{equation} \sigma_{1}=\frac{\alpha\beta h^{2}}{2}(1-\frac{1}{\tilde d }), ~~~~\sigma_{2}=\frac{\beta k^{2}}{2}(\frac{1}{\tilde d }-1) \label{917} \end{equation} The three-block p-branes solution for the Lagrangian with one differential form for various dimensions of the space-time was found in \cite{AV}. It contains previously known D=10 case \cite{Tseytlin,CM}. Equations of motion for the case of one form corresponds to equation of motion for ansatz (\ref{12}), (\ref{18}) and (\ref{1999}) for the dual action (\ref{11}) when $\alpha$= $\beta$ and $q=\tilde{d}$. Note that the metric (\ref{12}) describes also the solution for the action with the form $F_{q+1}$ replaced by its dual $F_{\tilde{q}+1}$ with $\tilde{q}+q+2=D$ and $\alpha \to \tilde{\alpha}=-\alpha$. One can also change two forms $F$ and $G$ to their dual version without changing the metric (\ref{12}). To illustrate our method on a simple example we first consider in Sect. 3 the simple case when in (\ref{011}) $d =q$ and one has only two blocks in the metric. Then in Sect. 4 we derive the solution (\ref{12}). In Sect. 5 we consider particular cases of the solution (\ref{12}) and obtain different known solutions. In Appendix the solution of the system of algebraic equations is given. \section{Two block solution} To illustrate the method of solution in this section we consider the simple case when the system of algebraic equations can be easily solved. Let us consider the action \begin{equation} I=\frac{1}{2\kappa ^{2}}\int d^{D}z\sqrt{-g} [R-\frac{1}{2}(\partial \phi) ^{2}- \frac{e^{-\alpha \phi}}{2(d+1)!}F^{2}_{d+1} -\frac{e^{\beta \phi}}{2(d+1)!}G^{2}_{d+1}) \label{31} \end{equation} The Einstein equations for the action (\ref{11}) read \begin{equation} R_{MN}-\frac{1}{2}g_{MN}R=T_{MN}, \label{19a} \end{equation} where the energy-momentum tensor has the form $$T_{MN}= \frac{1}{2} (\partial _{M}\phi \partial _{N}\phi - \frac{1}{2}g_{MN} (\partial \phi)^{2}) $$ $$ + \frac{1}{2d!}e^{-\alpha \phi} ( F_{MM_{1}...M_{d}}F_{N}^{M_{1}...M_{d}}- \frac{1}{2(d+1)}g_{MN}F^{2})+ $$ \begin{equation} \frac{1}{2d!}e^{\beta \phi} (G_{MM_{1}...M_{d}}G_{N}^{M_{1}...M_{d}}- \frac{1}{2(d+1)}g_{MN}G^{2}) \label{32} \end{equation} and one has also equations of motion \begin{equation} \partial _{M}(\sqrt{-g}e^{-\alpha \phi}F^{MM_{1}...M_{d}})=0 \label{33} \end{equation} \begin{equation} \partial _{M}(\sqrt{-g}e^{\beta \phi}G^{MM_{1}...M_{d}})=0 \label{34} \end{equation} \begin{equation} \partial _{M}(\sqrt{-g}g^{MN}\partial _{N }\phi ) + \frac{\alpha }{2(d+1)!}\sqrt{-g}e^{-\alpha \phi}F^{2}- \frac{\beta }{2(d+1)!}\sqrt{-g}e^{\beta \phi}G^{2}=0 \label{35} \end{equation} We use the following Ansatz \begin{equation} F=d{\cal A} , ~~~~~{\cal A}_{01...d-1}=\gamma _{1}e^{C_{1}(x)} \label{36} \end{equation} \begin{equation} G=d{\cal B} ,~~~~~ {\cal B}_{01...d-1}=\gamma _{2}e^{C_{2}(x)} \label{37} \end{equation} \begin{equation} ds^{2}=e^{2A(x)}\eta_{\alpha \beta} dy^{\alpha} dy^{\beta} +e^{2B(x)}dx^{i}dx^{i}, \label{38} \end{equation} $\alpha$, $\beta$ =0,1,...,d-1, $\eta_{\alpha \beta}$ is a flat Minkowski metric, $i,j$ =d,...,D. With the above Ansatz equations (\ref{19a}) are reduced to the following system of equations: \begin{eqnarray} (d-1)\Delta A+ (\tilde{d}+1)\Delta B + \frac{d(d-1)}{2}(\partial A)^{2} + \frac {\tilde{d}(\tilde{d}+1)}{2}(\partial B)^{2}+ (d-1)\tilde{d}(\partial A\partial B) = \nonumber \\ -\frac{\gamma _{1}^{2}}{4}e^{\alpha \phi -2dA +2c_{1}}(\partial c_{1})^{2}- \frac{\gamma _{2}^{2}}{4}e^{\alpha _{2}\phi -2dA +2c_{2}}(\partial c_{1})^{2}- \frac{1}{4}(\partial \phi)^{2} \label{39} \end{eqnarray} and $$ -d(\partial _{m}\partial _{n} A +\partial_{m}A \partial _{n}A ) -{\tilde d}(\partial _{m}\partial _{n}B -\partial _{m}B\partial _{n}B) +d(\partial _{m}A\partial _{n} B+\partial _{m}B\partial _{n} A)$$ $$ +\delta _{mn}[d\Delta A + \frac{d(d+1)}{2}(\partial A)^{2} +{\tilde d}\Delta B+ \frac{{\tilde d}({\tilde d}+1)}{2}(\partial B)^{2} +d({\tilde d}-1)(\partial A \partial B)]=$$ $$\frac{1}{2}[\partial _{m}\phi\partial_{n} \phi -\frac{1}{2}\delta_{mn} (\partial \phi)^{2}] -\frac{\gamma _{1}^{2}}{4}e^{\alpha \phi -2dA +2C_{1}} [\partial _{m}C_{1}\partial _{n}C_{1} -\frac{1}{2}\delta _{mn}(\partial C_{1})^{2}] $$ \begin{equation} -\frac{\gamma _{2}^{2}}{4}e^{-\beta \phi -2dA +2C_{2}} [-\partial _{m}C_{2}\partial _{n}C_{2} -\frac{1}{2}\delta _{mn}(\partial C_{2})^{2}] \label{310} \end{equation} Here $\tilde d=D-d-2$. Equations for antisymmetric fields are \begin{equation} \partial _{m}(e^{-\alpha \phi -dA +{\tilde d}B +c_{1}}\partial _{m}c_{1})=0 \label{311} \end{equation} \begin{equation} \partial _{m}(e^{\beta \phi -dA +{\tilde d}B +c_{1}}\partial _{m}c_{2})=0 \label{312} \end{equation} Equation of motion for dilaton $$\partial _{m}(e^{-dA +{\tilde d}B}\partial _{m}\phi ) -\frac{\alpha\gamma _{1}^{2}}{2}e^{-\alpha \phi -2dA +2C_{1}} (\partial _{m}C_{1})^{2}$$ \begin{equation} +\frac{\beta \gamma _{1}^{2}}{2}e^{\beta \phi -2dA +2C_{2}} (\partial _{m}C_{2})^{2} =0 \label{31m} \end{equation} We impose the following relations: \begin{equation} dA +{\tilde d}B =0, \label{314} \end{equation} \begin{equation} -\alpha \phi -2dA +2C_{1}=0, \label{315} \end{equation} \begin{equation} \beta \phi -2dA +2C_{2}=0. \label{316} \end{equation} Under these conditions equations (\ref{311}), (\ref{312}) and (\ref{31m}) take the following forms, respectively, \begin{equation} \partial _{m}(e^{-C_{1}}\partial _{m}C_{1})=0, \label{317} \end{equation} \begin{equation} \partial _{m}(e^{-C_{2}}\partial _{m}C_{2})=0, \label{318} \end{equation} \begin{equation} \Delta \phi -\frac{\alpha \gamma _{1}^{2}}{2}(\partial _{m} C _{1})^{2} +\frac{\beta \gamma _{2}^{2}}{2}(\partial _{m} C _{2})^{2}=0 \label{319} \end{equation} >From equations (\ref{317}) , (\ref{318}) and (\ref{319}) we get \begin{equation} \Delta C _{1} =(\partial C_{1})^{2}, \label{320} \end{equation} \begin{equation} \Delta C _{2} =(\partial C_{2})^{2} \label{321} \end{equation} and \begin{equation} \phi=\varphi _{1}C_{1}+\varphi _{2}C_{2} \label{322} \end{equation} with \begin{equation} \varphi _{1}=\frac{\alpha \gamma ^{2}}{2} \label{323} \end{equation} \begin{equation} \varphi _{2}=-\frac{\beta\gamma ^{2}}{2} \label{324} \end{equation} Equations (\ref{315}) and (\ref{316}) give \begin{equation} A=a_{1}C_{1}+a_{2}C_{2}, \label{325} \end{equation} where \begin{equation} a_{1}=\frac{\beta }{d(\alpha +\beta )},~~~ a_{2}=\frac{\alpha }{d(\alpha +\beta )}, \label{326} \end{equation} and \begin{equation} \phi=\varphi _{1}C_{1}+\varphi _{2}C_{2} \label{532} \end{equation} where \begin{equation} \varphi _{1}=\frac{2}{\alpha +\beta },~~~ \label{327} \end{equation} \begin{equation} \varphi _{2}=-\frac{2}{\alpha +\beta }. \label{328} \end{equation} Comparing (\ref{323}) with (\ref{327}) and (\ref{324}) with (\ref{328}) we conclude that \begin{equation} \gamma _{1}^{2}=\frac{4}{\alpha (\alpha +\beta )},~~~ \gamma _{2}^{2}=\frac{4}{\beta (\alpha +\beta )}. \label{329} \end{equation} Let us now consider equation (\ref{310}). Since $C_{1}$ and $C_{2}$ are two independent functions we have to have that the terms with $\partial _{m}C_{1}\partial _{n}C_{2}$ vanish, i.e. we have to impose the condition \begin{equation} \alpha \beta=\frac{2d{\tilde d}}{d+{\tilde d}}. \label{330} \end{equation} Therefore, the Ansatz (\ref{37}) is consistent with the metric (\ref{38}) only under condition (\ref{330}). Straitforward calculations show that for $A$, and $\phi$ given by (\ref{325})-(\ref{326}) and (\ref{532}) in terms of two independent functions $C_{1}$ and $C_{2}$, satisfying equations (\ref{320}) and (\ref{321}), equations (\ref{39}) (\ref{310}) are satisfied if we impose the condition (\ref{330}). Let us write the metric in terms of two harmonic functions: \begin{equation} H_{1}=-\ln C_{1},~~~ H_{2}=-\ln C_{2} \label{331} \end{equation} We finally get $$ds^{2}=H_{1}^{-\frac{2\beta }{d(\alpha +\beta )}} H_{2}^{-\frac{2\alpha }{d(\alpha +\beta )}} \eta_{\alpha \beta} dy^{\alpha} dy^{\beta} $$ \begin{equation} +H_{1}^{\frac{2\beta }{{\tilde d}(\alpha +\beta )}} H_{2}^{\frac{2\alpha }{{\tilde d} (\alpha +\beta )}}dx^{i}dx^{i} \label{332} \end{equation} \renewcommand{\theequation}{\thesection.\arabic{equation}} \section{Three block solution} \setcounter{equation}{0} In this secction we consider equations of motion for the action (\ref{11}). The Einstein equations for the action (\ref{11}) read \begin{equation} R_{MN}-\frac{1}{2}g_{MN}R=T_{MN}, \label{19} \end{equation} where the energy-momentum tensor is $$ T_{MN}= \frac{1}{2} (\partial _{M}\phi \partial _{N}\phi - \frac{1}{2}g_{MN} (\partial \phi)^{2}) $$ $$ + \frac{1}{2q!}e^{-\alpha \phi}(F_{MM_{1}...M_{q}}F_{N}^{M_{1}...M_{q}}- \frac{1}{2(q+1)}g_{MN}F^{2}) $$ \begin{equation} + \frac{1}{2{\tilde d}!}e^{-\alpha \phi}(G_{MM_{1}...M_{{\tilde d}}} G_{N}^{M_{1}...M_{{\tilde d}}}- \frac{1}{2({\tilde d}+1)}g_{MN}G^{2}) \label{110} \end{equation} The equation of motion for the antisymmetric fields are \begin{equation} \partial _{M}(\sqrt{-g}e^{-\alpha \phi}F^{MM_{1}...M_{q}})=0, \label{111a} \end{equation} \begin{equation} \partial _{M}(\sqrt{-g}e^{-\beta \phi}G^{MM_{1}...M_{{\tilde d}}})=0, \label{111} \end{equation} and one has the Bianchi identity \begin{equation} \epsilon ^{M_{1}...M_{q+2}}\partial_{M_{1}}F_{M_{2}...M_{q+2}}=0. \label{112} \end{equation} \begin{equation} \epsilon ^{M_{1}...M_{{\tilde d}+2}} \partial_{M_{1}}G_{M_{2}...M_{{\tilde d}+2}}=0. \label{1129} \end{equation} The equation of motion for the dilaton is \begin{equation} \partial _{M}(\sqrt{-g}g^{MN}\partial _{N }\phi ) + \frac{\alpha}{2(q+1)!}\sqrt{-g}e^{-\alpha \phi}F^{2} + \frac{\beta }{2({\tilde d}+1)!}\sqrt{-g}e^{-\beta \phi}G^{2} =0. \label{113} \end{equation} We shall solve equations (\ref{19})-(\ref{113}) by using the following Ansatz for the metric \begin{equation} ds^{2}=e^{2A(x)}\eta_{\mu \nu} dy^{\mu} dy^{\nu}+ e^{2F(x)}\delta_{nm}dz^{n}dz^{m}+ e^{2B(x)}\delta_{\alpha\beta}dx^{\alpha}dx^{\beta}, \label{114} \end{equation} where $\mu$, $\nu$ = 0,...,q-1, $\eta_{\mu\nu}$ is a flat Minkowski metric, $m,n$=$1,2,...,r$ and $\alpha,\beta$ =$1,...,{\tilde d}+2$. Here $A$, $B$ and $C$ are functions on $x$; $\delta_{nm}$ and $\delta_{\alpha\beta}$ are Kronecker symbols. Non-vanishing components of the differential forms are \begin{equation} {\cal A}_{\mu_{1}...\mu_{q}}=h{\epsilon_{\mu_{1}...\mu_{q}}}e^{C(x)}, ~F=d{\cal A} \label{115} \end{equation} \begin{equation} G^{\alpha_{1}...\alpha _{{\tilde d}+1}}=\frac{1}{\sqrt{-g}} k e^{\beta \phi}\epsilon ^{\alpha_{1}...\alpha_{{\tilde d}+1}\gamma} \partial _{\gamma} e^{\chi} , \label{116} \end{equation} where $h$ and $k$ are constants. The left hand side of the Einstein equations for the metric (\ref{114}) read $$R_{\mu\nu}-\frac{1}{2}g_{\mu\nu}R=\eta_{\mu\nu}e^{2(A-B)}[(q-1)\Delta A + ({\tilde d}+1)\Delta B + r\Delta F $$ $$+\frac{q(q-1)}{2}(\partial A)^{2}+ \frac{r(r+1)}{2}(\partial F)^{2}+\frac{{\tilde d}({\tilde d}+1)}{2}(\partial B)^{2}$$ \begin{equation} +{\tilde d}(q-1)(\partial A\partial B)+r(q-1)(\partial A\partial F) +r{\tilde d}(\partial F\partial B)], \label{175} \end{equation} $$R_{mn}-\frac{1}{2}g_{mn}R=\delta_{mn}e^{2(F-B)}[q\Delta A+ ({\tilde d}+1)\Delta B + (r-1)\Delta F $$ $$ +\frac{q(q+1)}{2}(\partial A)^{2}+ \frac{r(r-1)}{2}(\partial F)^{2}+\frac{{\tilde d}({\tilde d}+1)}{2}(\partial B)^{2}$$ \begin{equation} +{\tilde d}q(\partial A\partial B)+q(r-1)(\partial A\partial F) +{\tilde d}(r-1)(\partial F\partial B)], \label{176} \end{equation} $$R_{\alpha \beta} -\frac{1}{2} g_{\alpha \beta} R =-q\partial_{\alpha}\partial_{\beta} A- {\tilde d}\partial_{\alpha}\partial_{\beta} B- r\partial_{\alpha}\partial_{\beta} F$$ $$-q\partial_{\alpha} A\partial_{\beta}A+ {\tilde d}\partial_{\alpha} B\partial_{\beta}B- r\partial_{\alpha} F\partial_{\beta} F$$ $$+q(\partial_{\alpha} A\partial_{\beta} B+ \partial_{\alpha} B\partial_{\beta} A)+ r(\partial_{\alpha} B\partial_{\beta} F+ \partial_{\alpha} F\partial_{\beta} B)$$ $$+\delta_{\alpha\beta}[q\Delta A+ {\tilde d}\Delta B + r\Delta F + \frac{q(q+1)}{2}(\partial A)^{2}+ \frac{r(r+1)}{2}(\partial F)^{2}+ \frac{{\tilde d}({\tilde d}-1)}{2}(\partial B)^{2}$$ \begin{equation} q({\tilde d}-1)(\partial A\partial B)+ qr(\partial A\partial F) +r({\tilde d}-1)(\partial F\partial B)]. \label{177} \end{equation} For more details see \cite{AV}. Now one reduces the $(\mu\nu)$-components of (\ref{19}) to the equation \begin{eqnarray} (q-1)\Delta A+ ({\tilde d}+1)\Delta B +r\Delta F \nonumber \\ +\frac{q(q-1)}{2}(\partial A)^{2} + \frac{r(r+1)}{2}(\partial F)^{2}+ \frac {{\tilde d}({\tilde d}+1)}{2}(\partial B)^{2} \nonumber \\ +{\tilde d}(q-1)(\partial A\partial B) + r(q-1)(\partial A\partial F) + r{\tilde d}(\partial B\partial F) = \nonumber \\ -\frac{1}{4}(\partial \phi)^{2} -\frac{{h}^{2}}{4}(\partial C)^{2}e^{-\alpha\phi-2qA+2C} -\frac{k^{2}}{4}(\partial \chi)^{2} e^{2{\tilde d}B+\beta\phi+2\chi} , \label{117} \end{eqnarray} $(nm)$-components of (\ref{19}) to the following equation: $$q\Delta A+({\tilde d}+1)\Delta B+(r-1)\Delta F$$ $$+\frac{q(q+1)}{2}(\partial A)^{2}+ \frac{{\tilde d}({\tilde d}+1)}{2}(\partial B)^{2}+ \frac{r(r-1)}{2}(\partial F)^{2}$$ $$+q{\tilde d}(\partial A\partial B)+q(r-1)(\partial A\partial F)+ {\tilde d}(r-1)(\partial B\partial F)=$$ \begin{equation} -\frac{1}{4}(\partial \phi)^{2}+\frac{h^{2}}{4}(\partial C)^{2}e^{-\alpha\phi-2qA+2C}- \frac{k^{2}}{4}(\partial \chi)^{2}e^{2{\tilde d}B+\beta\phi+2\chi}, \label{118} \end{equation} and $(\alpha\beta)$-components to the equation: $$ -q\partial_{\alpha}\partial_{\beta} A- {\tilde d}\partial_{\alpha}\partial_{\beta} B- r\partial_{\alpha}\partial_{\beta} F $$ $$-q\partial_{\alpha} A\partial_{\beta} A + {\tilde d}\partial_{\alpha} B\partial_{\beta} B - r\partial_{\alpha} F\partial_{\beta} F + q(\partial_{\alpha} A\partial_{\beta} B +\partial_{\alpha} B\partial_{\beta} A)$$ $$+r(\partial_{\alpha} B\partial_{\beta} F + \partial_{\alpha} F\partial_{\beta} B) + \delta_{\alpha\beta}[q\Delta A+{\tilde d}\Delta B+ r\Delta F$$ $$+\frac{{\tilde d}({\tilde d}+1)}{2}(\partial A)^{2} + \frac{r(r+1)}{2}(\partial F)^{2} $$ $$ +\frac{{\tilde d}({\tilde d}-1)}{2}(\partial B)^{2} + q({\tilde d}-1)(\partial A\partial B) + r({\tilde d}-1)(\partial F\partial B )+ qr(\partial A\partial F)] $$ $$ = \frac{1}{2}[\partial_{\alpha} \phi\partial_{\beta} \phi - \frac{1}{2}\delta_{\alpha\beta}(\partial\phi)^{2}] - \frac{h^{2}}{2}e^{-\alpha\phi-2qA+2C}[\partial_{\alpha} C\partial_{\beta} C - \frac{\delta_{\alpha\beta}}{2}(\partial C)^{2}]$$ \begin{equation} -\frac{k^{2}}{2} e^{2{\tilde d}B+\beta\phi+ 2\chi}[\partial_{\alpha} \chi\partial_{\beta} \chi -\frac{\delta_{\alpha\beta}}{2}(\partial \chi)^{2}], \label{119} \end{equation} where we use notations $( \partial A\partial B)=\partial_{\alpha} A\partial_{\alpha} B$ and $D=q+r+{\tilde d}+2=d+{\tilde d}+2$. The equations of motion (\ref{111}) for a part of components of the antisymmetric field are identically satisfied and for the other part they are reduced to a simple equation: \begin{equation} \partial _{\alpha}(e^{-\alpha \phi -2qA +C}\partial _{\alpha}C)=0. \label{120} \end{equation} For $\alpha$-components of the antisymmetric field we also have the Bianchi identity: \begin{equation} \partial_{\alpha}(e^{\alpha\phi + 2Bq + \chi }\partial_{\alpha}\chi)=0.= \label{121} \end{equation} The equation of motion for the dilaton has the form $$\partial _{\alpha}(e^{qA +{\tilde d} B + Fr}\partial _{\alpha}\phi ) -\frac{\alpha h^{2}}{2}e^{-\alpha \phi -qA + qB +rF +2C}(\partial _{\alpha} C)^{2} $$ \begin{equation} +\frac{\beta k ^{2}}{2}e^{\beta \phi +2{\tilde d}B+2\chi} (\partial _{\alpha}\chi)^{2} =0. \label{122} \end{equation} We have to solve the system of equations (\ref{117})-(\ref{122}) for unknown functions$~~~~$ $ A,B,F,C,\chi,\phi$. We shall express $ A,B,F$ and $\phi $ in terms of two functions $C$ and $\chi$. In order to get rid of exponents in (\ref{117})-(\ref{122}) we impose the following relations: \begin{equation} qA + rF +{\tilde d}B =0, \label{123} \end{equation} \begin{equation} 2\chi +2{\tilde d}B +\beta \phi =0, \label{124} \end{equation} \begin{equation} 2C - 2qA -\alpha\phi = 0. \label{125} \end{equation} Under these conditions equations (\ref{120}),(\ref{121}) and (\ref{122}) will have the following forms, respectively, \begin{equation} \partial _{\alpha}(e^{-C}\partial _{\alpha}C)=0,~~~~~ \partial _{\alpha}(e^{-\chi}\partial _{\alpha}\chi)=0, \label{126} \end{equation} \begin{equation} \Delta \phi +\frac{\beta k ^{2}}{2}(\partial _{\alpha} \chi )^{2}- \frac{\alpha h^{2}}{2}(\partial _{\alpha} C )^{2}=0. \label{127} \end{equation} One rewrites (\ref{126}) as \begin{equation} \Delta C =(\partial C)^{2},~~~~~ \Delta \chi =(\partial \chi)^{2}. \label{128} \end{equation} Therefore (\ref{127}) will have the form \begin{equation} \Delta \phi +\frac{\beta k^{2}}{2}\Delta \chi - \frac{\alpha h^{2}}{2}\Delta C =0. \label{129} \end{equation} >From (\ref{129}) it is natural to set \begin{equation} \phi =\phi_{1}C + \phi_{2}\chi, \label{130} \end{equation} where \begin{equation} \phi_{1}=\frac{\alpha h^{2}}{2},~~~ \phi_{2}=-\frac{\beta k^{2}}{2}. \label{131} \end{equation} >From equations (\ref{123}), (\ref{124}) and (\ref{125}) it follows that $A$, $B$ and $F$ can be presented as linear combinations of functions $C$ and $\chi$: \begin{equation} A=a_{1}C + a_{2}\chi , \label{132} \end{equation} \begin{equation} B=b_{1}C + b_{2}\chi , \label{133} \end{equation} \begin{equation} F=f_{1}C + f_{2}\chi , \label{134} \end{equation} where \begin{equation} a_{1}=\frac{4-\alpha^{2}h^{2}}{4q}, ~~~ a_{2} = \frac{\alpha\beta k^{2}}{4q}, \label{135} \end{equation} \begin{equation} b_{1} = -\frac{\alpha\beta h^{2}}{4{\tilde d}},~~~ b_{2} =\frac{\beta ^{2}k^{2}-4}{4{\tilde d}}, \label{136} \end{equation} \begin{equation} f_{1} = \frac{\alpha^{2}h^{2} +\alpha \beta h^{2}-4}{4r},~~~ f_{2} = \frac{4-\alpha \beta k^{2}-\beta^{2} k^{2}}{4r}. \label{137} \end{equation} Let us substitute expressions (\ref{130}),(\ref{132})-(\ref{134}) for $\phi,A,B,F$ into (\ref{117})-(\ref{119}). We get relations containing bilinear forms over derivatives on $ C$ and $\chi$ . We assume that the coefficients in front of these bilinear forms vanish. Then we get the system of twelve quartic equations which is presented and solved in Appendix. The system has a solution only if $\alpha$ and $\beta$ satisfy the relation \begin{equation} \alpha \beta=\frac{2q{\tilde d}}{q+r+{\tilde d}} \label{81m} \end{equation} In this case $h$ and $k$ are given by the formulae \begin{equation} h=\pm\sqrt{\frac{4(q+r+{\tilde d})} {\alpha ^{2}(q+r+{\tilde d})+2q({\tilde d}+r)}}, \label{819} \end{equation} \begin{equation} k=\pm\frac{2\alpha (q+r+{\tilde d})}{\sqrt{{\tilde d} [2\alpha ^{2}(q+r)(q+r+{\tilde d})+4q^{2}{\tilde d}]}} \label{195} \end{equation} To summarize, the action (\ref{11}) has the solution of the form (\ref{12}) expressed in terms of two harmonic functions $H_{1}$ and $H_{2}$ if the parameters in the action are related by (\ref{193}) and the parameters in the Ansatz $h$ and $k$ are given by (\ref{819}),(\ref{195}). \section {Discussion and Conclusion} Let us discuss different particular cases of the solution (\ref{12}). There is the relation (\ref{15}) between parameters $\alpha$ and $\beta$ in the action (\ref{011}). As a result the action corresponds to the bosonic part of a supergravity theory only in some dimensions. If $D=4$ and $ q=d=1$ then one can take $\alpha=\beta=1$ and the action corresponds to the $SO(4)$ version of $N=4$ supergravity. The solution (\ref{12}) takes the form \begin{equation} ds^{2}=-H_{1}^{-1}H_{2}^{-1}dt^{2}+H_{1}H_{2}dx^{\alpha}dx^{\alpha} \label{195a} \end{equation} This supersymmetric solution has been obtained in \cite{kallosh}. If $\alpha=\beta$ and $q=\tilde d$ then one has the solution \begin{equation} ds^{2}=H_{1}^{\frac{2}{D-2}}H_{2}^{\frac{2(D-q-2)}{q(D-2)}} [(H_{1}H_{2})^{-\frac{2}{q}} \eta_{\mu \nu} dy^{\mu} dy^{\nu}+ H_{2}^{-\frac{2}{q}}dz^{m}dz^{m}+ dx^{\alpha}dx^{\alpha}], \label{713} \end{equation} This solution was obtained in \cite{AV}. It contains as a particular case for $d=10,~q=2$ the known solution \cite{TS,CM,CH} $$ ds^{2}=H_{1}^{-\frac{3}{4}}H_{2}^{-\frac{1}{4}} (-dy_{0}^{2}+dy_{1}^{2}) $$ \begin{equation} + H_{1}^{\frac{1}{4}}H_{2}^{-\frac{1}{4}} (dz_{2}^{2}+dz_{3}^{2}+dz_{4}^{2}+dz_{5}^{2})+ H_{1}^{\frac{1}{4}}H_{2}^{\frac{3}{4}} (dx_{6}^{2}+dx_{7}^{2}+dx_{8}^{2}+dx_{9}^{2}). \label{66m} \end{equation} This solution has been used in the D-brane derivation of the black hole entropy \cite{SV,CM}. Note however that the solution (\ref{66m}) corresponds to the action (\ref{011}) with the $3$-form $F_{3}$ and the $7$-form $G_{7}$. To conclude, a rather general three-block solution of the action (\ref{011}) has been constructed. It contains as particular cases many known solutions. However the solution is not general enough to include some known multi-block solutions. Further work is needed to understand better the structure hidden behind the multi-block p-brane solutions. \section*{Acknowlegments} This work is supported by an operating grant from the Natural Sciences and Engineering Research Council of Canada. I.A. and I.V. thank the Department of Physics for the kind hospitality during their stay at Simon Fraser University. I.V. is partially supported by the RFFI grant 9600312. \section {Appendix} We obtain the following system of algebraic equations $$-a_{1}+b_{1}+\frac{q(q-1)}{2}a_{1}^{2}+\frac{r(r+1)}{2}f_{1}^{2}+ \frac{{\tilde d}({\tilde d}+1)}{2}b_{1}^{2}$$ \begin{equation} +{\tilde d}(q-1)a_{1}b_{1}+r(q-1)a_{1}f_{1}+r{\tilde d}f_{1}b_{1}+ \frac{\phi_{1}^{2}}{4}+ \frac{h^{2}}{4}=0; \label{178} \end{equation} $$-a_{2}+b_{2}+\frac{q(q-1)}{2}a_{2}^{2} + \frac{r(r+1)}{2}f_{2}^{2} + \frac{{\tilde d}({\tilde d}+1)}{2}b_{2}^{2}$$ \begin{equation} +{\tilde d}(q-1)a_{2}b_{2}+r(q-1)a_{2}f_{2}+r{\tilde d}b_{2}f_{2} +\frac{\phi_{2}^{2}}{4}+\frac{k^{2}}{4}=0; \label{179} \end{equation} $$q(q-1)a_{1}a_{2} +r(r+1)f_{1}f_{2}+{\tilde d}({\tilde d}+1)b_{1}b_{2}$$ \begin{equation} +{\tilde d}(q-1)(a_{1}b_{2}+a_{2}b_{1})+r(q-1)(a_{2}f_{1}+a_{1}f_{2})+ r{\tilde d}(f_{1}b_{2} +f_{2}b_{1})+\frac{\phi_{1}\phi_{2}}{2}=0; \label{180} \end{equation} $$-f_{1}+b_{1}+\frac{q(q+1)}{2}a_{1}^{2}+\frac{r(r-1)}{2}f_{1}^{2}+ \frac{{\tilde d}({\tilde d}+1)}{2}b_{1}^{2}$$ \begin{equation} +q {\tilde d} a_{1}b_{1}+q(r-1)a_{1}f_{1}+ {\tilde d}(r-1)f_{1}b_{1}+\frac{\phi_{1}^{2}}{4} -\frac{h^{2}}{4}=0; \label{181} \end{equation} $$-f_{2}+b_{2}+\frac{q(q+1)}{2}a_{2}^{2} + \frac{r(r-1)}{2}f_{2}^{2} + \frac{{\tilde d}({\tilde d}+1)}{2}b_{2}^{2}+$$ \begin{equation} +q {\tilde d} a_{2}b_{2}+q(r-1)a_{2}f_{2}+{\tilde d}(r-1)b_{2}f_{2} +\frac{\phi_{2}^{2}}{4}+\frac{k^{2}}{4}=0; \label{182} \end{equation} $$q(q+1)a_{1}a_{2} +r(r-1)f_{1}f_{2}+{\tilde d}({\tilde d}+1)b_{1}b_{2}$$ \begin{equation} +q{\tilde d}(a_{1}b_{2}+a_{2}b_{1})+q(r-1)(a_{2}f_{1}+a_{1}f_{2}) +{\tilde d}(r-1)(f_{1}b_{2} +f_{2}b_{1})+\frac{\phi_{1}\phi_{2}}{2}=0; \label{183} \end{equation} $$-qa_{1}^{2}-rf_{1}^{2}+ {\tilde d} b_{1}^{2}+$$ \begin{equation} +2qa_{1}b_{1}+2rf_{1}b_{1}-\frac{\phi_{1}^{2}}{2}+ \frac{h^{2}}{2}=0; \label{184} \end{equation} $$-qa_{2}^{2} -rf_{2}^{2} + {\tilde d}b_{2}^{2}$$ \begin{equation} +2qa_{2}b_{2}+2rb_{2}f_{2} -\frac{\phi_{2}^{2}}{2}+\frac{k^{2}}{2}=0; \label{185} \end{equation} $$-qa_{1}a_{2} -rf_{1}f_{2}+{\tilde d}b_{1}b_{2}$$ \begin{equation} +q(a_{1}b_{2}+a_{2}b_{1})+r(f_{1}b_{2} +f_{2}b_{1})-\frac{\phi_{1}\phi_{2}}{2}=0; \label{186} \end{equation} $$\frac{q(q+1)}{2}a_{1}^{2}+\frac{r(r+1)}{2}f_{1}^{2}+ \frac{{\tilde d}({\tilde d}-1)}{2}b_{1}^{2}$$ \begin{equation} +q({\tilde d}-1)a_{1}b_{1}+r({\tilde d}-1)b_{1}f_{1}+ rqf_{1}a_{1}+\frac{\phi_{1}^{2}}{4}- \frac{h^{2}}{4}=0; \label{187} \end{equation} $$\frac{q(q+1)}{2}a_{2}^{2} + \frac{r(r+1)}{2}f_{2}^{2} + \frac{{\tilde d}({\tilde d}-1)}{2}b_{2}^{2}$$ \begin{equation} +q({\tilde d}-1)a_{2}b_{2}+r({\tilde d}-1)a_{2}f_{2}+rqb_{2}f_{2} +\frac{\phi_{2}^{2}}{4}-\frac{k^{2}}{4}=0; \label{188} \end{equation} $$q(q+1)a_{1}a_{2} +r(r+1)f_{1}f_{2}+{\tilde d}({\tilde d}-1)b_{1}b_{2}+ q({\tilde d}-1)(a_{1}b_{2}+a_{2}b_{1}) $$ \begin{equation} +r({\tilde d}-1)(b_{2}f_{1}+b_{1}f_{2})+rq(f_{1}a_{2} +f_{2}a_{1})+\frac{\phi_{1}\phi_{2}}{2}=0. \label{188a} \end{equation} Now let us discuss the system of eq. (\ref{178})-(\ref{188a}). The action depends on the parameters $D,\alpha, \beta ,q,\tilde{d}$. We have used the parameter $r$ instead of $D$ which is defined from $$q+r+\tilde{d} +2=D$$ Therefore our action (\ref{11}) depends on five parameters $r,\alpha, \beta ,q,\tilde{d}$ We have also two parameters $h$ and $k$ in our Ansatz (\ref{115}),(\ref{116}). We substitute expressions (\ref{130}),(\ref{131}), (\ref{135})-(\ref{137}) into (\ref{178})-(\ref{188a}). Then we get the system of twelve quartic equations for seven unknown variables $r,\alpha, \beta ,q,\tilde{d},h$ and $k$. Using Maple V we found the solution of the system of equations (\ref{178})-(\ref{188a}). The solution has the form \begin{equation} k=\frac{\sqrt{2q(r+{\tilde d})h^{2}-4(q+r+{\tilde d})}} {\sqrt{{\tilde d}(qrh^{2}-2(q+r))}}, \label{190} \end{equation} \begin{equation} \alpha=\frac{\sqrt{4(q+r+{\tilde d})-2q(r+{\tilde d})h^{2}}} {h\sqrt{q+r+{\tilde d}}} \label{191} \end{equation} \begin{equation} \beta=\frac{2q{\tilde d}h}{\sqrt{(q+r+{\tilde d}) [4(q+r+{\tilde d})-2q(r+{\tilde d})h^{2}]}} \label{192} \end{equation} where $r,q,\tilde{d},h$ are arbitrary. Let us rewrite the solution in terms of parameters in the action. >From equations (\ref{191}) and (\ref{192}) follows that $h$ can be found from equations (\ref{178})-(\ref{188a}) only if $\alpha$ and $\beta$ are subjected to the relation \begin{equation} \alpha \beta=\frac{2q{\tilde d}}{q+r+{\tilde d}} \label{193} \end{equation} In this case $h^{2}$ and $k^{2}$ are given by the formulae \begin{equation} h^{2}=\frac{4(q+r+{\tilde d})} {\alpha ^{2}(q+r+{\tilde d})+2q({\tilde d}+r)}, \label{} \end{equation} \begin{equation} k^{2}=\frac{2\alpha^{2} (q+r+{\tilde d})^{2}} {{\tilde d}[\alpha ^{2}(q+r)(q+r+{\tilde d})+2q^{2}{\tilde d}]} \label{} \end{equation} \newpage
proofpile-arXiv_065-411
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} In recent years, new semiconductor heterostructures have attracted considerable interest. Multiple quantum well structures and superlattices of II--VI compounds are the subject of intensive study because of their interesting optical properties \cite{tersoff,flores,brasil,ichino}. With these structures, energy gaps ranging from the UV to IR are accessible \cite{brasil,ichino,pelhos}. In these systems the binary interfaces are usually lattice mismatched. This lattice mismatch modifies the band alignments, and hence modifies the device optical properties. In searching for the desired material parameters such as band gap, lattice matching to substrates, dielectric contact, carrier mobility, etc., a large number of materials, have been investigated. Recently, high--quality cubic--structured ternary and quaternary alloys have been proposed as appropriate materials for heterostructures \cite{brasil,% ichino}. Ternary alloys allow a certain control of the induced strain at the interface. The deep understanding of the physics of the interface is important for the detailed study of thermal, optical, and other properties of quantum--wells and superlattices. The electronic properties at solid--solid interfaces depend sometimes even on details of the interaction between the two atomic layers from the different materials in contact. Our work can be used as a starting point to analyze those details. These are responsible for the characteristics of interface reconstruction, thermodynamic properties, degree of intermixing, stress, compound formation, etc. In previous work, we have studied the electronic structure of the valence band for the (001)--surface of several II--VI wide band gap semiconductors \cite{prb50,prb51}, and different binary heterostructures \cite{infcs}. We have obtained the (001)--projected electronic structure for both, surfaces and binary interfaces using the known Surface Green's Function Matching (SGFM) method \cite{g-m}. In the (001)--surfaces in addition to the well known bulk bands and surface resonances, we have described three different structures in the valence band region, the so--called surface induced bulk states ($B_h,\ B_l$, and $B_s$). We have shown that these states owe their origin to the creation of the surface, that is, they depend on the surface through the boundary condition (the wave function has to be zero at the surface), but they are not surface resonances. They are {\it surface--induced bulk states} \cite{prb50,prb51}. Later we found that this kind of induced states appear at the interface domain as well. Therefore, more generally, we found that any frontier can induce these states. For that reason we have redefined them as frontier--induced semi-infinite medium (FISIM) states since they are not, strictly speaking, {\it bulk} (infinite medium) states. These FISIM states do not show dispersion as a function of the wave vector {\bf k} for the surfaces studied. This is theoretically and experimentally shown for the (001)--oriented CdTe surface \cite{prb50,prb51,niles,gawlik}. For the binary/binary (001)--oriented II--VI compound interfaces, in contrast, they show some clear dispersion \cite{infcs}. The interest of the present work is twofold. Firstly, we want to make practical use of our recently set found tight--binding Hamiltonians for the ternary alloys. They reproduce the known experimental change with the composition of the band gap and they can be further used in detailed studies of different physical problems as, for example, the dependence of the transport properties on composition in quantum well structures that avoid stress. The second interest of this work is the study of the evolution of the FISIM states from a non--dispersive character to a dispersive one as stress and different crystal composition enters into play. We show that, if we select a ternary alloy to produce little stress and change only slightly the composition, the FISIM states do exists on both sides of the interface but do not show as much dispersion. So the existence of the FISIM states is due to the existence of a frontier alone and the amount of dispersion is related to the existing stress at the interface and on the chemical character of the interface partner. We will present in this work the valence band of some II--VI (001)--binary/ternary alloy interfaces and we will concentrate in particular in the FISIM states. The method used is discussed in our previous work. Here we only summarize the relevant features of it in Section II for completeness; Section III is devoted to discuss our results. Finally, we give our conclusions in section IV. \section{The Method} To describe the interface between two semiconductor compounds, we make use of tight--binding Hamiltonians. The Green's function matching method takes into account the perturbation caused by the surface or interface exactly, at least in principle, and we can use the bulk tight--binding parameters (TBP) \cite{rafa,noguera,quintanar}. This does not mean that we are using the same TBP for the surface, or for the interface and the bulk. Their difference is taken into account through the matching of the Green's functions. We use the method in the form cast by Garc\'\i a--Moliner and Velasco \cite{g-m}. They make use of the transfer matrix approach first introduced by Falicov and Yndurain \cite{falicov}. This approach became very useful due to the quickly converging algorithms of L\'opez--Sancho {\it et al.} \cite{sancho} Following the suggestions of these authors, the algorithms for all transfer matrices needed to deal with these systems can be found in a straightforward way \cite{trieste}. The Green's function for the interface, $G_I$, is given by \cite{g-m}, \begin{equation} \label{infcs} G_I^{-1}=G_{s(A)}^{-1} + G_{s(B)}^{-1} - I_BH^iI_A - I_AH^iI_B, \end{equation} where $G_{s(A)}$ and $G_{s(B)}$ are the surface Green's function of medium A and B, respectively. $-{\cal I}_AH^i{\cal I}_B$ and $-{\cal I}_BH^i{\cal I}_A$ are the Hamiltonian matrices that describe the interaction between the two media. In our model these are $20\times20$ matrices, the input TBP for these matrices are the average for those of the two media. This is a reasonable approximation when both sides of the interface have the same crystallographic structure and we take the same basis of wave functions. The tight--binding Hamiltonians for the II--VI ternary alloys are described in detail in Ref. [18]. Briefly speaking, we have used the tight--binding method and, under certain conditions, the virtual crystal approximation to study the ternary alloys. We have included an empirical bowing parameter\ in the $s-$on site TBP of the substituted ion. This procedure gave us the correct behaviour of the band gap value with composition \cite{ternario}. More exactly for the TBP of the ternary alloy, we take \begin{equation} \overline E_{\alpha,\alpha'}(x)= x E_{\alpha,\alpha'}^{(1)} + (1-x)E^{(2)}_{\alpha,\alpha'}, \qquad \alpha,\ \alpha'=s,\ p^3,\ s^* \label{vca} \end{equation} for all but the $s-$on site TBP of the substituted ion. In eq. (\ref{vca}) $E^{(1,2)}_{\alpha,\alpha'}$ are the TBP for the compound 1 (2); $\alpha,\ \alpha'$ are the atomic orbitals used in the basis set. For the $s-$on site TBP of the substituted ion we use the following expression \begin{equation} \overline E_{s,\nu}(x,b_\nu)= \overline E_{s,\nu}(x) + x(1-x)b_\nu, \qquad \nu=a,\ c \end{equation} where $\overline E_{s,\nu}(x)$ is given by eq. (\ref{vca}) and $b_\nu$ is the empirical bowing parameter\ per each different substitution (anion--substitution $(a)$ or cation--one $(c)$). In Table \ref{ebp} we have the empirical bowing parameter s used in this work. We do not introduce any further parameter \cite{ternario}. From the knowledge of the Green's function, the local density of states can be calculated from its imaginary part integrating over the two--dimensional first Brillouin zone, the dispersion relations can be obtained from the poles of the real part. We have applied previously this formalism to surfaces \cite{prb50,prb51,rafa,noguera}, interfaces \cite{infcs,quintanar,rafa-prb} and superlattices \cite{rafa-moliner}. Now we present our results. \section{Results and discussion} This section is devoted to the discussion of the interface--valence band of the (001)--projected electronic band structure of II--VI binary/ternary alloy interfaces. We will present in this paper the (001)--CdTe/CdSe$_{.15}$Te$_{.85}$, (001)--CdTe/Zn$_{.17}$Cd$_{.83}$Te, (001)--ZnSe/ZnSe$_{.87}$Te$_{.13}$, and (001)--ZnSe/Zn$_{.85}$Cd$_{.15}$Se interfaces in detail. The interfaces studied have been chosen with a composition $(x)$ as to give a minimum stress. For the lattice parameter value of the materials considered see Table II. As we can see the induced stress is small, about 1\%. This magnitude of the induced stress allow us to ignore its effect in our calculation. The real bulk bands as well as the FISIM states, should lie very closely to our calculated ideal case. We adopt the same convention for the interface domain as in Ref. \cite{infcs}. That is to say, we consider nearest neighbors interactions in our bulk Hamiltonians and, as a consequence, four atomic layers as the interface domain, two belonging to medium A and two to medium B. To distinguish between the different atomic layers we will call each atomic layer by the medium its neighbors belong to. The atomic layer AA will be the second from the interface into medium A. AB will be the last atomic layer belonging to medium A and facing the first atomic layer of medium B and so on. So the four atomic layers that constitute the interface domain will be labeled AA, AB, BA, and BB. For the interfaces aligned along the (001) direction the two media are facing each other either through its anion or cation atomic layer. In the alloy case, we consider a pseudobinary compound so that the concept of anion and cation atomic layers remain meaningful. We will consider here only anion-anion interfaces but our results can be extented without difficulty to other kind of interfaces. We will project the interface electronic band structure on each atomic layer and we will see how the different states that we found for the free surface and for the binary/binary interface case change or disappear at the binary/ternary one. It is known that the common anion interfaces have small valence band--offset and the common cation ones have small conduction band--offset, both of the order of some meV \cite{ichino,pelhos,% duc}. In consequence, we will use the boundary condition that the top of the valence bands at the interface are aligned and choose this energy as our zero. Accordingly, the conduction band offset will be equal to the difference in the band gaps. The actual calculation of the band offset is still an open theoretical question that we do not want to address in this work \cite{infcs2}. As a general remark, the FISIM states are not Bloch states and therefore the {\bf k}--wave number is not expected to be a good quantum number. The existence of a frontier (surface or interface) breaks the symmetry. This does not actually mean that when the Schr\"odinger equation is solved for differents values of {\bf k} (the Hamiltonian depends explicitly on it) one should get the same eigenvalue. It is found, theoretically \cite{prb50,prb51} as well as experimentally \cite{niles,gawlik}, that the solution does not depend on {\bf k} for the case of a surface. In this case the boundary condition is that the wave function has to be zero at the surface boundary for any value of the derivative. It is the condition for an infinite potential barrier. For the interface it is not so. For the binary/binary case we got a solution that depends on the wave vector, {\bf k}, but we should not call it {\it dispersion} since it is not the behaviour with respect to a quantum number that we are looking at but rather with respect to a parameter. FISIM states are neither Bloch states nor surface states. They do exist in the semi--infinite medium space but they do not follow the infinite--medium symmetry of the crystal. So we have to look for a different physical reason of their {\bf k}--dependence. The first thing to notice is that the boundary condition is different. For an interface, the wave function has not to be zero, is has to be continuous together with the derivative. The boundary condition therefore will depend on {\bf k}. This is because the Hamiltonian describing the interaction depends on it and therefore the wave function that solves the Schr\"odinger equation does depend on it as well. For this reason its value and its derivative at the border will also depend on it. One does therefore, in general, expect a {\bf k}--dependence of the FISIM states eigenvalues for an interface. For a surface the boundary condition is always zero and on the contrary we do not expect a {\bf k}--dependence. In previous work,\cite{infcs} we have explored the behaviour of the FISIM states at binary/binary interfaces. These represent a strong change at the interface. In this work, we explore the existence and behaviour of the FISIM states at interfaces that do change slowly. Here, ternary alloys are chosen so as to minimize stress (same lattice constant in both sides) and the corresponding binary/ternary alloy interface FISIM states are obtained. Their {\bf k}--dependence as expected, is minimum. So, we can conclude that, in general, stress is responsible for the {\bf k}--dependence of the FISIM states. This is in agreement with the ideas developed above. Therefore, FISIM states are a consequence of the existence of a frontier and their {\bf k}--dependence is a result of the stress at it. Furthermore, we have obtained from this calculation two interface states in the valence band range for the CdTe--based interfaces and one interface state for the ZnSe--ones. Now we present the details for each interface. In Figs. 1--4, we show the electronic band structure of the valence band for the interfaces studied here, (001)--CdTe/CdTe$_{.85}$Se$_{.15}$, (001)--CdTe/Zn$_{.17}$Cd$_{.83}$Te, (001)--ZnSe/ZnSe$_{.87}$Te$_{.13}$, and (001)--ZnSe/Zn$_{.85}$Cd$_{.15}$Se. The dispersion relations are found from the poles (triangles in the figures) of the real part of the interface Green's function. The solid--lines are a guide to the eye. These are to be compared to the dispersion curves found for the bulk (infinite medium) case. The calculated eigenvalues for the FISIM states are denoted by stars, crosses and points; the dotted lines are intended only as a guide to the eye. We label the FISIM states as $B_{Ih},\ B_{Il}$, and $B_{Is}$. This convention follows the previous free (001)--surfaces study (see Refs. \cite{prb50,prb51}). The energy eigenvalues for all the calculated states are in Tables III and IV. \subsection{The (001)--CdTe/CdSe$_{.15}$Te$_{.85}$ interface} Fig. 1 shows the projected electronic structure of the valence band for this interface per atomic layer. From the figure is evident that we have obtained the general pattern of the projected band structure of the II--VI semiconductor surfaces \cite{prb50,prb51}. As we have commented above we will consider an anion--anion interface and we will aling the top of the valence band as our zero of energy. We have obtained that the heavy hole (hh) and light hole (lh) bands show more dispersion in the interface domain than in the semi--infinite medium. They are usually low in energy about 0.7 eV an 0.4 eV, respectively, in all the atomic layers. The spin--orbit band shows almost the same dispersion that in the semi--infinite medium, see Table III. As is pointed previously \cite{infcs}, the FISIM states $B_{Ih}$ and $B_{Il}$ in the interface domain do not mix with the hh and lh bands, as is observed in the (001)--surface case \cite{prb50,prb51}, see Fig. 1 and Table IV. The states are lower in energy than the lh band. These upper FISIM states show a slight dependence on {\bf k}, but in most of the cases it is less than 0.3 eV. In contrast, the $B_{Is}$ state follows the spin--orbit band as in the semi--infinite medium. In general, from the Fig. 1 we appreciate that the FISIM states show better behaviour than in the binary/binary interfaces \cite{infcs}. Moreover, in this energy interval we have obtained some states that we identify with {\it interface states} ($IS_1$ and $IS_2$, the dotted lines in Fig. 1, are a guide to the eye). The first one, at --1.3 eV, in $\Gamma$, shows notable dispersion and seems to disappear for {\bf k}--values near the $X$--point. The second state, with more noticely dispersion, appears at --1.9 eV in $\Gamma$ and reaches the $X-$point in --4.3 eV. However, as we do not know about experimental results in this system we can not give a complete comparison. We only predict the possibility of the existence of these interface states. \subsection{The (001)--CdTe/Zn$_{.17}$Cd$_{.83}$Te interface} This system shows almost the same pattern describe above. The calculated valence band electronic structure is presented in Fig. 2. From the Table III we observe that the hh and lh bulk bands show more dispersion that in the semi--infinite medium. In particular the electronic structure projected onto the Cd--atomic layer (Fig. 2a).) shows bigger dispersion for these bands, of about 0.8 and 0.5 eV, respectively, than the semi--infinite medium, see Table III. For the other atomic layers the projected electronic structure shows almost the same pattern all together: the hh and lh bands are 0.5 and 0.3 eV below in energy with respect to the bulk values, respectively. The spin--orbit band, however, in the interface seems to form a barrier of about 0.2 eV from the AA--atomic layer to the BB--atomic layer, see Table III. In general, the $B_{Ih},\ B_{Il}$ and the $B_{Is}$ FISIM states are lower in energy than the hh, lh, and spin--orbit bulk bands, respectively. In the same way that in the previous case, these FISIM states do not mix with the respective bulk bands at $X$, as in observed in the semi--infinite medium case \cite{prb50,prb51}. However, the FISIM states shows slight dependence on {\bf k}. As in the previous interface, we obtain two interface states in the present system, label $IS_1$ and $IS_2$ in Fig. 2. The $IS_1$ state, located at --1.3 eV in $\Gamma$, shows notable dispersion and reaches the $X-$point between the hh and lh bulk bands. The $IS_2$ state, with bigger dispersion than the previous one, is located in $\Gamma$ at --1.9 eV and reaches the $X-$point at the same values that the spin--orbit bulk band. \subsection{The (001)--ZnSe/ZnSe$_{.87}$Te$_{.13}$ interface} Fig. 3 shows the calculated electronic structure of the valence band for this interface. Opposite to the CdTe--based interfaces, discused above, the bulk bands and the FISIM calculated states for this system shows almost the same behaviour that in the semi--infinite medium. This observation goes for all the calculated bands but the spin--orbit band in the $\Gamma$ point, where we obtain, as in the previous case, a discontinuity from the AA--atomic layer to the BB--atomic layer. In this case the spin--orbit band seems to form a potential well, in $\Gamma$, of about 0.2 eV, see Table III. On the other hand, for this interface we obtain that the $B_{Ih},\ B_{Il}$, and $B_{Is}$ FISIM states mix with the hh, lh, and spin--orbit bulk bands, respectively, as is observed in the semi--infinite medium \cite{prb50,prb51}. In this sense this interface shows better behaviour than the other ones \cite{infcs}. Although, as previously, we have obtained an interface state for this system, $IS_1$. This interface state appears in $\Gamma$ at --1.7 eV, shows noticely dispersion and seems to disappear for {\bf k}--values near the $X-$point. However, the state appears notoriously in all the calculated atomic layers. \subsection{The (001)--ZnSe/Zn$_{.85}$Cd$_{.15}$Se interface} Finally, in the Fig. 4 we show our electronic structure calculated for this system. In the same way that the previous case, we obtain that all the calculated states, per atomic layer, for this interface are similar with the semi--infinite medium, see Tables III and IV. In addition to these states, we have an interface state, $IS_1$, located in $\Gamma$ at --1.7 eV and showing notable dispersion. The state do not appear for all the interval between $\Gamma-X$, it seems to disappear for {\bf k}--values near the $X-$point, as we have commented previously for the other interfaces. \section{Conclusions} In conclusion, we have calculated the electronic structure of the valence band of the II--VI binary/ternary alloy interfaces. We have used the tight--binding method and the surface Green's function matching method to obtain the electronic structure projected onto each atomic layer that constitutes the interface domain. For the ternary alloys we have used our tight--binding Hamiltonians described in previous work that give good account for the changes of the band gap with composition as obtained experimentally. Our parametrization includes an empirical bowing parameter\ for the ``$s$'' on--site tight--binding parameter of the substituted ion and we use the known virtual crystal approximation for the rest of them. The systems were chosen here so that stress can be ignored for the particular value of the compositional variable. The calculated valence band electronic structure of these interfaces show bulk bands with similar dispersion as for the semi--infinite medium (a system with a surface). The FISIM states observed in the (001)--oriented surfaces and binary/binary interfaces appear also in this case and show an intermediately strong {\bf k}--dependence as compare to the previous ones. In the interface domain the calculated states, both the bulk bands and the FISIM states, have a composition that is a combination of the corresponding states of the two media forming the interface. It is interesting to note further that we have obtained for the binary/ternary alloy case two interface states for the CdTe--based heterostructures and one interface state for the ZnSe--ones that do not show for the binary/binary interfaces at least in the energy interval that we have considered. We will consider the binary/quaternary and the ternary/quaternary alloy interfaces in future work.
proofpile-arXiv_065-412
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} A photographic plate consists of an emulsion which separates grains of AgCl (or similar) molecules. Let me present a simplified discussion of the detection of a single photon. The photon may dissociate an AgCl molecule. The Cl escapes, whereas the Ag radical starts, upon photographic development, a chemical reaction which leaves a small, visible spot. Before detection the one-particle photon wave may pass through {\it both} parts of a double slit, see for instance ref.\cite{Bra}. The geometry of the device (distance between the slits etc.) can be chosen such that the photon is spread out over a region covering distances much larger than a single grain, which determines the size of the finally visible spot. The physics which causes the photon wave to collapse is not understood. Let us first recall where the Schr\"odinger equation leaves us. For simplicity we assume that each grain consists of precisely one AgCl molecule. We label the AgCl molecules on the plate by $i=1,...,n$ and assume that the relevant features of the photographic plate are described by products of AgCl molecule wave functions. Initially the state is \begin{equation} |\Psi\rangle = |\Psi_0\rangle = \prod_{i=1}^n |\psi_i^b\rangle , \ \langle \Psi_0 | \Psi_0 \rangle = 1, \label{Psi0} \end{equation} where the $|\psi_i^b\rangle$, $(i=1,...,n)$ indicate the {\it bound} AgCl molecules and the overlap between different molecules is neglected, {\it i.e.} $\langle \psi_i^b | \psi_j^b \rangle = \delta_{ij}$. Through interaction with the photon the state is transformed into \begin{equation} \label{causal} |\Psi\rangle = c_0 |\Psi_0\rangle + \sum_{j=1}^n c_j |\Psi_j\rangle ~~{\rm with}~~ |\Psi_j\rangle = |\psi_j^d\rangle \prod_{i\ne j} |\psi_i^b\rangle ,\ \langle \Psi_j | \Psi_j \rangle = 1\, . \label{Psi} \end{equation} Here $|\psi_j^d\rangle$, $(j=1,...,n)$ denotes a {\it dissociated} AgCl molecule. A quantum measurement is constituted by the fact that only one of the $|\Psi_j\rangle$, $(j=0,1,...,n)$ states survives, each with probability $P_j=|c_j|^2$, $\sum_{j=0}^n P_j = 1$. The probabilities $P_j$ are related to the photon wave function $\psi_{ph}$ by means of \begin{equation} P_j = const\ \int_{V_j} d^3x\ |\psi_{ph}|^2\, ,\ j=1,...,n\, . \label{probability} \end{equation} The constant does not depend on $j$, and $V_j$ is a cross-sectional volume corresponding to the $j$th molecule. Picking a branch $|\Psi_j\rangle$ becomes in this way interpreted as observing the photon at the position $V_j$ (for a perfect detector $c_0=0$). It is remarkable that in this process of measurement the photon becomes destroyed through spontaneous absorption by the dissociating molecule. To summarize, measurements perform wave function {\it reductions} by making decisions between alternatives proposed by the continuous, causal time evolution part of Quantum Theory (QT). In our example the reduction decides the location of the visible spot. Given the photon wave function $\psi_{ph}$, QT predicts probabilities for the reduction alternatives. Whereas time evolution from eqn.(\ref{Psi0}) to (\ref{Psi}) is described by the Schr\"odinger equation, this is not true (or at least controversial) for the measurement process. Decoherence theory, for an overview see \cite{Zurek}, tries to establish that conventional time evolution leads, in the situation of eqn.(\ref{Psi}), to $\langle\Psi_j | \Psi_k\rangle = 0$ for $j\ne k$, such that it becomes impossible to observe contradictions with interference effects predicted by the Schr\"odinger equation. In contrast to this, explicit collapse models predict deviations from conventional QT, see for instance \cite{Ch86}. So far, no such deviations have been measured. The Schr\"odinger equation, more precisely its applicable relativistic generalization $|\Psi (t)\rangle = \exp ( - H\, t)\, |\Psi (0)\rangle $, describes a continuous and causal time evolution and I shall use the notation Quantum Object (QO) for matter $|\Psi (t)\rangle$ as long as it exhibits this behavior. Concerning the measurement process, it seems to be widely believed that many body processes, involving $\gg 10^{10}$ particles \cite{Bra}, are responsible. In the presented example the photographic plate, possibly also the environment beyond, would be blamed. However, the ultimate collapse into one branch $|\Psi_j\rangle$ remains an unexplained property. No satisfactory derivation from the known properties of microscopic matter appears possible. Consequently, a search for hereto overlooked new, fundamental properties of matter is legitimate. The question arose whether hidden variables may exist which ensure local, continuous and causal time development for the entire system, including measurements. Bell \cite{Bell} turned this apparently philosophical question into physics by showing that all such local, realistic theories are measurably distinct from QT (Bell's inequalities). Subsequently, many experiments were performed and local, realistic theories are now convincingly excluded. For instance, the experiments of \cite{Aspect} found violations of Bell's inequalities for spacelike measurements on entangled quantum states. In such experiments one performs measurements at distinct locations, say $\vec{x}_1$ and $\vec{x}_2$, in time intervals small enough that any mutual influence through communication at or below the speed of light can be excluded. Results at $\vec{x}_1$ correlate with those at $\vec{x}_2$ (and vice versa) in a way that {\it excludes} an interpretation as a classical correlation, see \cite{Mermin} for a pedagogical presentation. Such effects underline the need for qualitatively new properties of matter, because they cannot be propagated through local, relativistic wave equations (which also govern the interaction with the environment). Also a consistent description of the space-time evolution of the quantum state vector $|\Psi\rangle$ under such measurements encounters difficulties. After accepting that a measurement at $(c\, t_1,\vec{x}_1)$ or $(c\, t_2,\vec{x}_2)$ interrupts the continuous, causal time evolution by a discontinuous jump, once faces the problem that Lorentz transformations can change the time ordering of spacelike events. In a recent paper \cite{Be98a} it has been shown that a spacetime picture for a physical state vector with relativistically covariant reduction exists. It may be summarized as follows: \begin{description} \item{(1)} Measurement are performed by detectors, which are part of the state vector, at localized spacetime positions $(c\, t_i,\vec{x}_i)$, $i=1,2,\dots\,$. \item{(2)} Discontinuous reductions of the state vector are defined on certain Lorentz covariant spacelike hypersurfaces, which in some neighbourhood of a detector include its backward light cone. \item{(3)} The thus defined measurements happen in some reduction order, which is {\it not} a time ordering with respect to a particular inertial frame. \end{description} Based on this scenario, I pursue in the present paper an approach which builds on the strengths of QT and tries to supplement it with new laws for reductions, such that the conventional rules for measurements (Born's probability interpretation) follow. These laws are supposed to act on the {\it microscopic} level, independently of whether macroscopic measurements are actually carried out or not. Typically, they will effect some interference phenomena. This implies observable consequences and makes their eventual existence a physical issue. On the other hand, we will see that rules can be designed in a way that most interference effects survive entirely, whereas those affected are only weakened in the sense of a decreased signal over background ratio (visibility). That makes such laws difficult to detect, as most experiments work with ensembles of particles and are happy to demonstrate a small signal over a large (subtracted) background. Fortunately, recent years have seen considerable improvements of experimental techniques, such that invoking experimental input may become feasible. The central idea of my approach is to {\it propose} that the Ability To Perform Reductions (ATPR) between alternatives proposed by QT is a hereto unidentified {\it elementary} property of microscopic matter. I shall use the notation Quantum Detector (QD) to denote microscopic as well as macroscopic matter acting in its ATPR. Matter gets such a dual character: As QO it follows the continuous, deterministic time evolution. As QD it has the ATPR and causes jumps in the wave function. The goal is to explain the functioning of actually existing macroscopic detectors from the properties of microscopic QDs. Within our hypothetical framework central questions are now: \begin{description} \item{(1)} Which conglomerates of matter constitute a QD? \item{(2)} Which are precisely the alternatives of QT standing up for reductions? \item{(3)} What are the rules according to which QDs make their reductions? \end{description} It is unlikely, that ultimate answers can be found without additional experimental guidance. But it is instructive to introduce a simple hypothesis which allows (a) to illustrate the possibilities and general direction of the approach and (b) focuses on questions about experimental input which, quite generally, may be crucial for achieving progress in the field. Let us return to the detection of a photon by a photographic plate. The simplest possibility is to attribute to each {\it single} AgCl molecule the ATPR about collapsing the photon wave function. As this is a spontaneous absorption, we get to the question asked in the title of this paper. We assume that each molecule acts independently when making its reductions and, by chance, the $j^{th}$ molecule makes its reduction first, ahead of the others. The alternative is to decay or to stay intact. Either choice causes a jump in the wave function $|\Psi\rangle$ of eqn.(\ref{causal}). Subsequently rules are given which seem (a) to be minimal and (b) consistent with observations. The collapse results are fixed by the rules of quantum mechanics. Namely, to be either ($f$ stands for final) \begin{equation} \label{collapse1} |\Psi\rangle \to |\Psi\rangle_f = {c_j\over \sqrt{P_j}}\, |\Psi_j\rangle ~~~{\rm with\ probability}~~ P_j = |c_j|^2 \end{equation} or \begin{equation} \label{collapse2} |\Psi\rangle \to |\Psi'\rangle = \sum_{k\ne j} c'_k\, |\Psi_k\rangle ,\ c'_k= {c_k \over \sqrt{1-P_j}}, ~~~{\rm with\ probability}~~ 1-P_j\, . \end{equation} The $k$-sum in eqn.(\ref{collapse2}) includes $k=0$, compare (\ref{causal}). The particular choice of the phase factors, $c_j/\sqrt{P_j}$ and $c_k/\sqrt{1-P_j}$, assumes that a decoherence process leads into alternative branches, whereas for collapse with incomplete decoherence the issue would have to be resolved by the collapse rules. Each molecule thus constitutes a QD. As their mutual distances are short and their relative motion is negligible, compared to the speed of light, we can ignore the relativistic complications discussed in \cite{Be98a} (the backward light-cone becomes an excellent approximation to ``instantaneous''). Equation (\ref{collapse1}) implies as final result a dark spot at the position of molecule $j$. By construction this happens with the correct probability $P_j$. As soon as the wave function (\ref{collapse1}) rules, the reduction is completed. This is different when molecule $j$ stays intact. Then the same rules (\ref{collapse1}) and (\ref{collapse2}), which collapse $|\Psi\rangle$ of eqn.(\ref{causal}), have now to be applied to the wave function $|\Psi'\rangle$ of eqn.(\ref{collapse2}). Assume, molecule $l$ (note $l\ne j$ as the branch $|\Psi_j\rangle$ does no longer exist) makes the next reduction. The transformation will be either \begin{equation} \label{collapse3} |\Psi'\rangle \to |\Psi\rangle_f = {c'_l\over \sqrt{P'_l}}\, |\Psi_l\rangle ~~~{\rm with\ probability}~~ P'_l=|c'_l|^2 \end{equation} or $$ |\Psi'\rangle \to |\Psi''\rangle = \sum_{k\ne j,l} c''_k\, |\Psi_k\rangle ,\ c''_k= {c'_k \over \sqrt{1-P'_l}}, ~~{\rm with\ probability}~~ 1-P'_l\, . $$ Putting equations (\ref{collapse2}) and (\ref{collapse3}) together, we obtain $$ |\Psi\rangle \to |\Psi\rangle_f = {c_l\over \sqrt{P_l}}\, |\Psi_l\rangle ~~~{\rm with\ probability}~~ P_l,$$ {\it i.e.} precisely the correct likelihood to find the dark spot at the position of molecule $l$. Continuing the procedure, it is easy to see that all probabilities come out right. Once a molecule $j$, $j=1,...,n$ has collapsed $|\Psi\rangle$ into the $|\Psi\rangle_f$ state of eqn.(4), the chemical reaction$^1$ \footnotetext[1]{{During the chemical reaction similar collapse processes may continue. Presently, they are not of interest to us, as our aim is to discuss the collapse of the incoming photon wave function, which has the special property of being transversally spread out over a macroscopic region.}} -- initiated by the corresponding branch of each molecule -- survives only in the neighbourhood of molecule $j$, where the visible spot will occur. Let $t=0$ be the time at which the photon hits the photographic plate. This time is well-defined as long as we can assume that the photon flight time over a distance of the relevant thickness of the photographic plate (for example $0.3~mm \Rightarrow \triangle t = 10^{-12}~s = 1~ps$) is much smaller than the typical collapse time. Let us denote by \begin{equation} \label{tauf} P^f(t)=(1-P_0)\, (1-e^{-t/\tau^f(t)}) \end{equation} the probability that, at time $t$, the (entire) system has decided about the location of the dark spot. Here $P_0=|c_0|^2$, see eqn.(\ref{causal}), is the probability that the system fails to detect. The r.h.s. of (\ref{tauf}) defines the system collapse time $\tau^f(t)$. If $\tau^f$ is constant, it is the mean time the system needs to make its reduction (with corresponding collapse probability density $(\tau^f)^{-1}\, \exp (-t/\tau^f)$). Let us assume that each molecule performs reductions on its own and that $\rho^c_j(t)$ is the likelihood per time unit that the $j^{th}$ molecule makes its reduction. We simplify the situation further and consider a $\rho^c_j(t)$ that does not depend on $j$ and is a step function: $\rho_j^c(t)=\rho^c\, \theta(t)$ with $\rho^c =$ constant. The corresponding one molecule collapse probability is \begin{equation} \label{tauc} p^c(t)=(1-e^{-t/\tau^c})\, \theta(t)\, , \end{equation} where $\tau^c=1/\rho^c$ is the mean collapse time of a single molecule. The molecules make their reductions in some sequential order. For our purposes the reductions process comes to a halt as soon as one molecule has decayed. Assume, $n^c(t)$ molecules made their reductions. Whatever values the $P_j$ in eqn.(\ref{collapse1}) take, $P^f(t) = (n^c(t)/ n)\, (1-P_0)$ is the probability that the collapse process has selected a definite location. As $n^c(t)=n\, p^c(t)$, we conclude \begin{equation} P^f(t)= (1-P_0)\, p^c(t)\, . \end{equation} With the approximations made the system collapse time $\tau^f$, defined by (\ref{tauf}), and the single molecule collapse time $\tau^c$, defined by (\ref{tauc}), are identical. Soon some arguments will be given that the constant $\tau^c$ should be regarded as upper bound of the system collapse time $\tau^f(t)$. Let us return to the central questions. In our discussion of detection of a spread-out photon, I assumed the following: (1) Each, single AgCl molecule may act as QD. (2) One alternative stands up for reduction: decaying (through absorbing the photon) or staying intact. (3) Each AgCl makes its reduction with a certain, constant likelihood per time unit: $\rho^c$. Ad (1): Assuming that a single AgCl molecule can act as QD reflects the attempt to introduce an ATPR as a fundamental property of microscopic matter. In our simplified discussion each AgCl molecule is separated from the others by the emulsion and causal interactions between them can be neglected. In a real photographic film only grains of AgCl molecules are separated. Causal interactions between an AgCl molecule and its neighbors within one grain cannot be neglected. Indeed, the initiated chemical process will spread out over the entire grain. If the collapse time is sufficiently large, competing (ultimately alternative) chemical processes would start to evolve in several grains. Under such circumstances, the definition of the QD should be extended to include each causally connected region of AgCl molecules. As a general rule, I find it attractive to conjecture that a conglomerate of matter which (in a reasonable approximation) can be treated as isolated QO can also be regarded as isolated QD. Neglecting the influence of most of the world is precisely how we get solutions out of QT. The hypothesis is, whenever this works well for a QO, this QO may also constitute a QD whose reduction probabilities are determined by its local quantum state, although this quantum state may participate in discontinuous, non-local transformations. Ad (2): The scenario, pursued now, is that the QT alternatives up for reductions have, quite generally, to do with absorption and emission of particles. Here I limit the discussion to the absorption and emission of photons, the process argued to be at the heart of every real, existing and functioning measurement device. Of course, other physical processes (like for instance in nuclear decay) should then be governed by similar rules. In essence: Ruled by not yet identified {\it stochastic} laws, superpositions of Fock space sectors with distinct particle numbers are conjectured to collapse into particle number eigenstate sectors. Ad (3): Our ATPR introduces an explicit arrow in time. This is attractive, because it is a matter of fact that such an arrow exists. The canonical conjugate variable to time is energy. Therefore, a frequency law which relates the collapse time to the difference $\triangle E$ in energy distribution between emerging branches is suggested \begin{equation} \label{tcguess} \tau^c\ =\ \tau^c (\triangle E)\, , \end{equation} where the energy difference is defined as the one experienced by the QD. For example, in case of our single AgCl molecule the difference between absorbing or not absorbing the photon is: $\triangle E = E_{\gamma}$, where $E_{\gamma}$ is the energy of the photon. The total energy is (in the same way as in QT) conserved in our approach. The introduced system has been chosen because of its popularity in QT text books in connection with the double slit effect. Instead of the photographic plate other measurement devices can be considered. For instance, the emergence of a track in a bubble chamber through ionization by an high energy particle allows a similar discussion. Here it is instructive to consider the emergence of such a track in monatomic dilute gas, say hydrogen. Assume that {\it one} incoming high energy particle has been split into two distinct transversally sharp rays$^1$, \footnotetext[1]{{Within our approach it might, however, happen that such a state collapses spontaneously, because the device which caused the split (and hence correlates with it) might act as a QD.}} each with 50\% probability content. At time $t=0$ the two rays may hit spacelike regions of hydrogen gas. Each ray builds up a column of half-ionized atoms. Let us focus on one of them, consisting of $n$ participating atoms. If one of the atoms of our column emits a photon by re-capturing an electron the relevant reduction has been made. A transformation of type (\ref{collapse1}) puts all atoms of the competing column into their unperturbed branches and the atoms of our column into their ionized branches. There is now some ambiguity about what should be considered a QD. Should each single atom (including the involved electrons) act as independent QD or should all atoms of the column together form one, single QD? In favor of the first viewpoint is that the gas is assumed to be dilute. Hence, the mutual influence through continuous, causal time evolution between the atoms is negligible. On the other hand, the high energy particle correlates all the atoms within the column (and, of course, also the other column): If one atom performs its reduction in favor of the ionized branch, all other are put there too. Assume the atoms act independently and the differences in energy distribution between their branches are $\triangle E_i$, ($i=1,...,n$), implying corresponding mean collapse times $\tau^c_i (\triangle E_i)$. The probability that none of them makes the reduction during the time interval $[0,t]$ becomes $$ q^c (t) = \prod_{i=1}^n \exp [-t/\tau^c_i (\triangle E_i) ] = \exp \left[ - \sum_{i=1}^n t/\tau^c_i (\triangle E_i) \right]\, .$$ On the other hand, if they all together form one single QD, this probability becomes $$ q^c (t) = \exp [-t/\tau^c(\triangle E)]\, ,~~{\rm where}~~ \triangle E = \sum_{i=1}^n \triangle E_i $$ is the total difference in energy distribution between the alternative, macroscopic branches. Remarkably, the $q^c(t)$ of the last two equations agree, when the law for the mean collapse time is \begin{equation} \label{berg} \tau^c (\triangle E)\ =\ {b\, \hbar \over \triangle E}\, , \end{equation} where $b$ is a dimensionless constant. (Correspondingly, $\tau^c_i =b\, \hbar/\triangle E_i$, $i=1,..,n$, of course.) Equation (\ref{berg}) has phenomenologically attractive features. The first one is that the indicated ambiguity is rendered irrelevant. Another is that the collapse time becomes large for small energy differences. Especially, superpositions of states degenerate in energy will not collapse. In this context {\it a measurement device is now an apparatus which speeds up the collapse by increasing the difference in energy distribution between quantum branches.} Before the distinct branches become macroscopically visible, the energy difference becomes so large that collapse happens with (practical) certainty. Are there observable consequences beyond standard QT? Reduction by an AgCl molecule destroys the possibility of interference of the branches (\ref{collapse1}) and (\ref{collapse2}) of the wave function (\ref{Psi}). In particular, this does still hold for the case of a single molecule $(n=1)$. But it appears unlikely that anyone will, in the near future, measure interference effects between AgCl$+\gamma$ and Ag$+$Cl. Hence, there is no contradiction. In addition, it should be noted that our mechanism leaves the most commonly observed interference effects intact: Namely, all those which rely on the wave character of particles in a Fock space sector with fixed particle number. This includes photon or other particle waves passing through double slits and so on. Neutron interferometry which relies on hyperfine level splitting would, in principle, be suppressed. However, the energy differences are small such that observable effect are unlikely. Larger energy differences are achieved in atomic beam spectroscopy. Ramsey fringes have been observed from interference of branches which differ by photon quanta with energy in the $eV$ range. Figure~1 depicts the interaction geometry of Bord\'e \cite{Bo84}, for a recent review see~\cite{St97}. An atomic beam of two level systems $(E_0 < E_1)$ interacts with two counterpropagating sets of a traveling laser wave. The laser frequency is tuned to the energy difference $\triangle E=E_1-E_0$, such that induced absorption/emission processes take place at each of the four interaction zones. The laser intensity is adjusted such that at each interaction zone an incoming partial wave is (further) split into two equally strong parts, $|\psi_0,m_0\rangle$ and $|\psi_1,m_1\rangle$. Here $|\psi_0\rangle$ denotes an atom in its incoming state, $|\psi_1\rangle$ an excited atom and $m_0$, $m_1$ are the numbers of photon moments transferred. Examples are indicated in the figure. The process leaves us with $2^n$ partial waves after the $n^{th}$ interaction zone, $n=1,2,3,4$. Of the final sixteen partial waves $4\cdot 2=8$ interfere under detuning of the laser frequency. Positions and directions of those eight partial waves are along the four lines, indicated after the last interaction zone of figure~1. The interference can be made visible by monitoring the decay luminosity $I$ of the excited states $|\psi_1\rangle$ after the last interaction zone. The contrast or visibility is defined by \begin{equation} \label{contrast} K = {I_{\max} - I_{\min} \over I_{\max} + I_{min}}\, , \end{equation} where $I_{\max}$ and $I_{\min}$ are maximum and minimum of the measured luminosities. Eight of the final sixteen partial waves are in excited states and four of them interfere in two pairs, $|\psi_1,-1\rangle$ and $|\psi_1,1\rangle$ of the figure. Hence, the optimal contrast for the Bord\'e geometry is \begin{equation} \label{Kopt} K_{opt} = {(4-4) + (8-0) \over (4+4) + (8+0)} = 0.5\, . \end{equation} This result is found by normalizing (in arbitrary units) the average luminosity of each excited partial wave to one. Four decoherent branches contribute then $I_{\max}=I_{\min}=4$, whereas the other four excited partial waves contribute $I_{\min}=0$ and $I_{\max}=8$ (for $I_{\min}$ they annihilate one another and in the other extreme they amplify). According to our hypothesis, integer photon numbers get restored with a collapse time $\tau^c=\tau^c(\triangle E)$. If this happens in range~1 of figure~1, the interference effect becomes entirely destroyed. The likelihood for it to happen is $p^c=1-q^c$, where $q^{c}=\exp (-t_D/\tau^c)$ and $t_D$ is the time an atom stays in range~1. In range~2 each of the two $|\psi_1,1\rangle$ partial waves has borrowed $1/4$ of a photon from the laser beam. To get a unique collapse description, we invoke a minimality assumption: The splitting of the atom has to be constructed with the minimal number of photons possible. The assumption seems to be natural, because two photons with the same quantum numbers cannot be distinguished. It follows that the system can collapse either into the two $|\psi_1,1\rangle$ partial waves or into the two $|\psi_0,0\rangle$ partial waves. Neither collapse has observable consequences, because the interference effects of the upper part and lower part of figure~1 are not distinguished by measuring the decay luminosity. In range~3 two split photons (distinct momenta) get involved: One mediates collapse between $|\psi_1,1\rangle$ and $|\psi_0,2\rangle$, the other between $|\psi_0,0\rangle$ and $\psi_1,1\rangle$. These two collapse processes are supposed to act independently. Each destroys, if it happens, half of the interference effect. The probability for both of them to happen is $(p^c)^2$ (using that $t_D$ is identical in range~3 and~1) and the probability that one of them (excluding both) happens is $p'^c=1-(q^c)^2-(p^c)^2$. Putting things together, the optimal contrast becomes \begin{equation} \label{Kcopt} K_{opt}^c=16^{-1}[8-8\,p^c-4\,q^c\,p'^c-8\,q^c\,(p^c)^2]= 0.5\, \exp [ -2\, t_D/\tau^c (\triangle E)]\, . \end{equation} Experiments performed at the Physikalisch-Technische Bundesanstalt (PTB) Braunschweig rely on the $^3P_1$--$^1S_0$ transition of $^{40}Ca$ which has $\lambda = 657.46$~nm, {\it i.e.} $\triangle E =1.886\, eV$. The best contrast achieved \cite{Riehle} is approximately $K=0.2$ with $t_D=21.6\cdot 10^{-6}\,s$. The actual experiments are performed using pulsed laser beams applied to laser cooled atoms in a magneto-optical trap. The times $t_D$ and $t_d$, corresponding to the distances $D$ and $d$ of figure~1, are then the times between the laser pulses, see~\cite{St97} for details. Relying on the PTB result we obtain the estimate $$ \tau^c_{\min}\, (1.89\, eV) = 2\,t_D\, /\, \ln (5/2) = 47\cdot 10^{-6}\,s < \tau^c\, (1.89\, eV) $$ which translates (\ref{berg}) into \begin{equation} \label{bmin} b_{\min} = 1.35\cdot 10^{11} < b\, . \end{equation} That the constant $b$ has to be large is no surprise, as the action $b\hbar$ marks the transition from quantum to classical physics. The bound (\ref{bmin}) can easily be improved by estimating conventional effects which contribute to diminishing the contrast $K$. Beyond, a direct measurement of a non-zero $\tau^c$ requires that all other effects can convincingly be controlled and that still a gap between the estimated and measured contrast remains. Such an analysis goes beyond the scope of the present paper. Here, I am content with establishing firm, but crude, bounds on $b$. Finally, in this paper, I derive an upper bound on $b$. Avalanche photodiodes are the up-to-date devices for achieving measurements in short time intervals, as needed for spacelike measurements \cite{Aspect}. Time resolutions down to $20\,ps$ FWHM are achieved, see \cite{Co96} for a recent review. The energy consumption is sharply peaked in these short intervals (order of watts), but does not translate into an immediate estimate of the collapse time. The reason is that collapse at some later time may lead to indistinguishable results. Claiming differently includes the task of disproving popular decoherence ideas \cite{Zurek}. Nevertheless, there is an easy way to estimate upper bounds by analysis of actually working measurement devices: Our approach makes only sense when the reduction process does keep up with the {\it sustained} performance of every real, existing measurement device. Then, the energy dissipation of such a device yields immediately an upper bound on $b$. Ref.\cite{Co96} gives on p.1964 the example of a photo avalanche diode which operates at $10^5\,cps$ and has a mean power dissipation of $4\,mW$. This translates into an energy consumption of about $2.5\cdot 10^{11}\,eV$ per count, {\it i.e.} $$ \tau^c\, (2.5\cdot 10^{11}\, eV) < \tau^c_{\max}\,(2.5\cdot 10^{11}\, eV) = 10^{-5}\, s\, ,$$ which implies \begin{equation} \label{bmax} b < b_{\max} = 3.8\cdot 10^{21}\, . \end{equation} Equation (\ref{bmin}) and (\ref{bmax}) leave a wide range open. An analysis of existing experiments should allow to narrow things down by a least a few orders of magnitude. Here the emphasize is on quoting save, instead of sophisticated, bounds. Even this has caused some efforts, the reason simply being that experimentalists do not focus on the information needed. In conclusion, we have discussed the possibility of attributing to microscopic matter the ability to perform wave function reductions. It is of interest to improve the bounds $b_{\min}$ (\ref{bmin}) and $b_{\max}$ (\ref{bmax}) for the collapse time $\tau^c (\triangle E)$ of equation~(\ref{berg}). From this viewpoint, I would like to argue in favor of a paradigm shift concerning QT experiments. It is no longer of central interest to demonstrate the existence of one or another exotic interference effect. We know, they are there. Most interesting is to control that interference happens for every single, participating particle. This puts the focus on experiments with high visibility. If one could convincingly demonstrate that particles occasionally skip participation in an interference pattern, such a results could pave a major inroad towards understanding of the measurement process. The aim of pushing experiments towards optimal visibility is of interest in itself. Independent of its validity, the introduced collapse scenario provides an interesting classification pattern for such results: The achieved lower bounds $b_{\min}$ should be compiled. Concerning $b_{\max}$, one is lead to minimizing the energy dissipation of measurement devices under sustained performance. Again, this is a goal of interest in itself. \medskip \noindent {\bf Acknowledgements:} I would like to thank Dr. Wolfgang Beirl for his interest and useful discussions. \medskip
proofpile-arXiv_065-413
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\subsection{Simulation Results in $2$ Dimensions} The critical transition in the shape of the hysteresis loop is observed in the simulation, and expected from the renormalization group \cite{Dahmen1,Dahmen2}, in $3$, $4$, and $5$ dimensions. We also found that the upper critical dimension, at and above which the mean field exponents become correct, is six. Furthermore, in one dimension, we expect that in a thermodynamic system, with an unbounded distribution of random fields, there will be no infinite avalanche for $R > 0$. That will be so because if there is any randomness, there will be a spin in the linear chain that will have the ``right'' value for its random field to stop the first avalanche. For a bounded distribution of random fields, the scaling behavior near the transition will not be universal\ \cite{Dahmen1}; instead, it will depend on the exact shape of the tails of the distribution of random fields. Then, the question that remains is: what happens in two dimensions? From the simulation and a few arguments that we are about to show, we conjecture that the two dimensional exponents will have the values: $\tau + \sigma\beta\delta =2$, $\tau=3/2$, $1/\nu=0$, and $\sigma\nu =1/2$. (The other exponents (except $z$) can be found from exponent relations\ \cite{Dahmen1,Dahmen3} using these values.) The ``arguments'' are as follows. It is quite possible that two is the {\it lower critical dimension} (LCD) for our system. At the lower critical dimension, the critical exponents are often ratios of small integers, and it is often possible to derive exact solutions. Since the geometry in $2$ dimensions allows for at most one system spanning avalanche, the ``breakdown of hyperscaling'' exponent $\theta$ (see section IV B) must be zero, and the hyperscaling relation\ \cite{Dahmen1,Dahmen3} is restored: \begin{equation} {1 \over \sigma\nu} = d - {\beta \over \nu} \label{hyper_2d_equ1} \end{equation} We know that this relation is violated in $4$ and $5$ dimensions, and is probably violated in $3$ dimensions. In dimensions above two, the hyperscaling relation is modified by the exponent $\theta$ which gives a measure of the number of spanning avalanches near the critical transition, as a function of the system size. Figure\ \ref{span_2d_fig} shows the number of spanning avalanches in $2$ dimensions for several system sizes. Notice that, as assumed, there is not more than one spanning avalanche in the system. \begin{figure} \centerline{ \psfig{figure=Figures/Span_aval_paper_2d_THESIS.ps,width=3truein} } \caption[Spanning avalanches in $2$ dimensions] {{\bf Number of spanning avalanches in $2$ dimensions} as a function of disorder $R$, for several system sizes. The data points are averages between as little as $10$ to as many as $2200$ random field configurations. Some typical error bars near the center of the curves are shown; error bars are smaller toward the ends. Note that there is no more than one spanning avalanche. \label{span_2d_fig}} \end{figure} We use two more arguments to derive the critical exponents. In $2$ dimensions, we find that the avalanches ``look'' compact (figure\ \ref{compact_2d_fig}). (The avalanches in $3$ dimensions are not compact (figure\ \ref{notcompact_3d_fig}).) This implies that $1/\sigma\nu = d = 2$, which leads to $\beta/\nu=0$ from equation\ (\ref{hyper_2d_equ1}). Furthermore, it is often the case that in the lower critical dimension, the Harris criterion\cite{Dahmen1} \begin{equation} {\nu \over \beta\delta} \geq {2 \over d} \label{Harris_eqn} \end{equation} becomes saturated (an equality); so in $2$ dimensions we expect $\beta\delta/\nu=1$. From this and the previous result, the exponent which gives the decay in space of the avalanche correlation function \begin{equation} \eta = 2 + {\beta \over \nu} - {\beta\delta \over \nu} \label{eta_eqn} \end{equation} (see references\ \cite{Dahmen1,Dahmen3} for the derivation of all the exponent relations) becomes equal to $\eta =1$. Since at the LCD the correlation length typically diverges exponentially as the critical point is approached, we expect $\nu \rightarrow \infty$, and $\beta$ can be finite. Using the exponent relation\ \cite{Dahmen1,Dahmen3}: \begin{equation} \tau-2=\sigma\beta(1-\delta), \label{hyper_2d_equ3a} \end{equation} we further find that $\tau=3/2$ and $\tau+\sigma\beta\delta=2$. \begin{figure} \centerline{ \psfig{figure=Figures/2D_R.8_span_grey_THESIS.ps,width=3truein} } \caption[Simulation of a $400^2$ spin system at $R=0.8$] {Simulation in $2$ dimensions of a $400^2$ spin system at $R=0.8$. The figure shows the configuration of the system after a spanning avalanche has just occurred (grey region). The dark area corresponds to spins that have not yet flipped, while the white area are spins that have flipped earlier. Notice that the spanning avalanche (grey area) seems compact. \label{compact_2d_fig}} \end{figure} We must mention that our firm conjectures about the exponents in two dimensions must be contrasted with our lack of knowledge about the proper scaling forms. As mentioned above, at the LCD the correlation exponent $\nu$ typically diverges, although some combinations of critical exponents stay finite (hence $\sigma\nu = 1/2$). Those which diverge and those which go to zero usually must be replaced by exponents and logs, respectively. We have used three different RG-scaling {\it ans\"atze} to model the data in two dimensions. (1)~We used the traditional scaling form $\xi \sim |R_c-R|^{-\nu}$, deriving $\nu = 5.3\pm 1.4$ and $R_c = 0.54\pm 0.04$. These collapses worked as well as any, but the large value for $\nu$ (and larger value still for $1/\sigma = 10\pm 2$) makes one suspicious. (2)~We used a scaling form suggested by Bray and Moore\cite{BrayMoore} in the context of the equilibrium thermal random field Ising model at the LCD, where $R_c=0$: if they assume that $R$ is a marginal direction, then by symmetry the flows must start with $R^3$, leading to $\xi \sim e^{(\tilde{a}/|R_c-R|^2)} \equiv e^{(\tilde{a}/R^2)}$. This form has the fewest free parameters, and most of the collapses were about as good as the others (except notably for the finite-size scaling of the moments of the avalanche size distribution, which did not collapse well once spanning avalanches became common). (3)~We developed another possible scaling form, based on a finite $R_c$ and $R$ marginal, which generically has a quadratic flow under coarse-graining: here $\xi \sim e^{(\tilde{b}/|R_c-R|)}$. We find $R_c=0.42\pm0.04$. The rational behind these three forms is shown in appendix A. The results from data collapses in two dimensions were obtained from measurements of the spanning avalanches, the second moments of the avalanche size distribution, the integrated avalanche size distributions, and the avalanche correlations. The magnetization curves are also obtained from the simulation, but as in the higher dimensions, the scaling region is small (around $H_c$ and $M_c$), and the collapses do not define the exponents well. \begin{figure} \centerline{ \psfig{figure=Figures/Sites_1_grey_THESIS.ps,width=3truein} } \caption[Largest avalanche in the hysteresis loop in a $40^3$ system, near the critical point] {Largest avalanche occurring in the hysteresis loop in a $40^3$ spins system near the critical point. The avalanche is not compact. \label{notcompact_3d_fig}} \end{figure} Measurements that require the knowledge of the critical randomness are the binned avalanche size distribution from which we extract the exponents $\tau$ and $\beta\delta$, the critical magnetic field $H_c$, and the avalanche time measurement which gives the exponent $z$. These measurements were not obtained {\it at} the critical disorder because $R_c$ is not well defined as was mentioned above, and because for low disorders (less than $0.71$ for a $7000^2$ system), the system flips in one infinite avalanche, and such measurements are therefore not possible. We have nevertheless estimated the values of some of these exponents and of $H_c$, from data obtained at a larger disorder (where there is no spanning avalanche). From the avalanche size distribution binned in $H$ at $R=0.71$ and $L=7000$, and the magnetization curves, we find that the critical field $H_c$ is around $1.32$. A straight line fit through the data agrees with a possible value of $\tau=3/2$ (the conjectured value). From the time distribution of avalanche sizes for a system of $30000^2$ spins, at $R=0.65$, we measured (from a straight linear fit) the exponent $\sigma\nu z$ to be $0.64$. The other exponents were obtained from scaling collapses as follows. Figure\ \ref{s2_2d_fig}a shows the second moments of the avalanche size distribution for several system sizes. The collapses using the three different scaling forms are shown in figures\ \ref{s2_2d_fig}(b-d). The first one (figure\ \ref{s2_2d_fig}b) is: \begin{equation} {\langle S^2 \rangle}_{int} \sim L^{-(\tau+\sigma\beta\delta-3)/\sigma\nu}\ {\check {\cal S}}_{int}^{(2)}(L\ |r|^{\nu}) \label{2d_equ01} \end{equation} which is the kind of scaling form used in $3$, $4$, and $5$ dimensions. This form assumes $\xi \sim |r|^{-\nu}$. The exponents are $(\tau + \sigma\beta\delta-3)/\sigma\nu = -1.9$ and $\nu=5.25$, and $r=(R_c - R)/R$ with $R_c=0.54$. The second scaling form (figure\ \ref{s2_2d_fig}c) is: \begin{equation} {\langle S^2 \rangle}_{int} \sim L^{-(\tau+\sigma\beta\delta-3)/\sigma\nu}\ {\bar {\cal S}}_{int}^{(2)}(L\ e^{-\tilde{a}/|R_c-R|^2}) \label{2d_equ02} \end{equation} which is obtained from $\xi \sim e^{(\tilde{a}/|R_c-R|^2)}$. The values of the exponents and parameters are: $(\tau + \sigma\beta\delta-3)/\sigma\nu = -1.9$, $\tilde{a} = 3.4$ ($\tilde{a}$ is not universal), and $R_c \equiv 0$ (by assumption; see previous paragraph). Notice that this collapse is not as good as the other two; a better collapse is obtained with $R=0.15$ and $\tilde{a}=2.0$. If this is the correct scaling form and $R_c=0$, this discrepancy can be due to finite size effects. The third scaling form is (figure\ \ref{s2_2d_fig}d): \begin{equation} {\langle S^2 \rangle}_{int} \sim L^{-(\tau+\sigma\beta\delta-3)/\sigma\nu}\ {\hat {\cal S}}_{int}^{(2)}(L\ e^{-\tilde{b}/|R_c-R|}) \label{2d_equ03} \end{equation} which is obtained from $\xi \sim e^{(\tilde{b}/|R_c-R|)}$. The values of the exponents and parameters are: $(\tau + \sigma\beta\delta-3)/\sigma\nu = -1.9$, $\tilde{b} = 2.05$ ($\tilde{b}$ is also non-universal), and $R_c= 0.42$. As it is clear from the last three figures, collapses with these different scaling forms are comparable. Notice that the exponent $(\tau+\sigma\beta\delta -3)/\sigma\nu$ is the same for the three collapses, but that $1/\nu$ is zero for the last two (by assumption) while it is $0.19$ for the first collapse. Let's now look at the collapses of the integrated avalanche size distribution curves, which are not finite size scaling measurements. \begin{figure} \centerline{ \psfig{figure=Figures/Non_Span_s2_d2_paper_THESIS.ps,width=3truein} } \nobreak \centerline{ \psfig{figure=Figures/Non_Span_s2_d2_collapse_s2_paper_new_THESIS.ps,width=3truein} } \nobreak \centerline{ \psfig{figure=Figures/Non_Span_s2_d2_collapse_exp_n2_paper_THESIS.ps,width=3truein} } \nobreak \centerline{ \psfig{figure=Figures/Non_Span_s2_d2_collapse_exp_n1_paper_THESIS.ps,width=3truein} } \caption[Second moments of the avalanche size distribution in $2$ dimensions] {(a) {\bf Second moments of the avalanche size distribution in $2$ dimensions,} integrated over the external field $H$, for several system sizes. The data points are averages over up to $2200$ random field configurations. Error bars are smaller than shown for larger disorders. (b), (c), and (d) Scaling collapses of the second moments of the avalanche size distribution in $2$ dimensions, integrated over the field $H$. The curves that are collapsed are of size: $50^2$, $100^2$, $300^2$, $500^2$, $1000^2$, $3000^2$, $5000^2$, $7000^2$, and $30000^2$. See text for the scaling forms, and the values of the exponents and parameters. \label{s2_2d_fig}} \end{figure} Figure \ref{aval_2d_fig}a shows the integrated avalanche size distribution curves for a $7000^2$ spin system, at several values of the disorder $R$. Earlier, in figure\ \ref{bump_345fig}, we saw the fit to the scaling collapse of such curves, done using the same scaling form as in $3$, $4$, and $5$ dimensions: \begin{equation} D_{int}(S,R)\ \sim\ S^{-(\tau + \sigma\beta\delta)}\ {\bar {\cal D}}^{(int)}_{-} (S^{\sigma} |r|) \label{2d_equ1} \end{equation} (The $-$ sign indicates that the collapsed curves are for $r<0$, ie. $R>R_c$.) However, $S^\sigma |r|$ might not be the appropriate scaling argument in $2$ dimensions. First, from figure\ \ref{bump_345fig}, the scaling curve in $2$ dimensions differs dramatically from the scaling curves in higher dimensions for small arguments $X=S^\sigma |r|$. The mean field scaling function $\bar{\cal D}_{-}^{(int)}(X)$ is a polynomial for small $X$, and we expected (and found) a similar behavior in $5$, $4$ and $3$ dimensions (but notice that the scaling function in $3$ dimensions is starting to look like the curve in $2$ dimensions for small $X$). In $2$ dimensions, if we collapse our data (figure\ \ref{aval_2d_fig}b) using the scaling form: \begin{equation} D_{int}(S,R)\ \sim\ S^{-(\tau + \sigma\beta\delta)}\ {\cal D}^{(int)}_{-} (S|r|^{1/\sigma}) \label{2d_equ2} \end{equation} with $\tau + \sigma\beta\delta=2.04$, $1/\sigma=10$, and $r=(R_c - R)/R$, we find that the scaling function for small $\tilde X=S|r|^{1/\sigma}$ looks linear with power one! This might imply that the scaling function ${\cal D}^{(int)}_{-}(S|r|^{1/\sigma})$ (eqn.\ (\ref{2d_equ2})) is the one that is analytic for small arguments in $2$ dimensions. \begin{figure} \centerline{ \psfig{figure=Figures/Aval_histo_d2_L7000_paper_THESIS.ps,width=3truein} } \nobreak \centerline{ \psfig{figure=Figures/Aval_histo_d2_L7000_collapse_paper_THESIS.ps,width=3truein} } \nobreak \centerline{ \psfig{figure=Figures/Aval_histo_d2_L7000_collapse_exp_n2_paper_THESIS.ps,width=3truein} } \nobreak \centerline{ \psfig{figure=Figures/Aval_histo_d2_L7000_collapse_exp_n1_paper_THESIS.ps,width=3truein} } \caption[Integrated avalanche size distribution curves in $2$ dimensions] {(a) {\bf Integrated avalanche size distribution} curves for several disorders {\bf in $2$ dimensions,} at the system size $L=7000$. The curves are averages over $10$ to $20$ random field configurations, and have been smoothed. (b), (c), and (d) Scaling collapses of the data from (a) using the three scaling forms and the exponents from the text. The collapsed curves have disorders: $0.72$, $0.74$, $0.77$, and $0.80$. The straight grey line in each of the plots has a slope of one. \label{aval_2d_fig}} \end{figure} Second, we conjectured above that the values for $\sigma$ and $1/\nu$ are probably zero in $2$ dimensions, and that only the combination $\sigma\nu$ is finite ($\sigma\nu$ $=$ $1/2$). It follows that, for the other two scaling forms we use, the arguments of the scaling function should be $S e^{-\tilde{a}/\sigma\nu|R-R_c|^2}$ and $S e^{-\tilde{b}/\sigma\nu|R-R_c|}$, and {\it not} $S^\sigma e^{-\tilde{a}/\nu|R-R_c|^2}$ and $S^\sigma e^{-\tilde{b}/\nu|R-R_c|}$ respectively. This is analogous to using $S|r|^{1/\sigma}$ in the scaling form\ (\ref{2d_equ2}). We should mention here that both equation (\ref{2d_equ1}) and equation (\ref{2d_equ2}) give the same scaling exponents $\tau+\sigma\beta\delta$ and $\sigma$, and that in all our scaling collapses we have assumed that the same scaling argument is valid for small and large $\tilde X$ (and in between). This in general, does not have to be true. \begin{figure} \centerline{ \psfig{figure=Figures/Aval_histo_2d_L30000_R0.650_ave_fit_paper_THESIS.ps,width=3truein} } \caption[Linear fits to an integrated avalanche size distribution curve in $2$ dimensions] { {\bf Integrated avalanche size distribution curve in $2$ dimensions}, for a system of $30000^2$ spins, at $R=0.650$. Shown are two linear fits to the data: one for small sizes and the other for large sizes. The slope for the fit at small $S$ is $0.90$. The fit was done for sizes in the range $[10,250]$. The slope differs by less than $5\%$ when the range is changed ($S$ is never larger than $400$ though) . The slope for the fit at large $S$ is $1.78$. The slope differs by less than $2\%$ when the range is changed ($S$ is never smaller than $10000$). The conjectured value for $\tau+\sigma\beta\delta$ is $2$ which is different from $1.78$. This is similar to the behavior we saw in $3$, $4$, and $5$ dimensions. On the other hand, for small sizes we expect the exponent $\tau+\sigma\beta\delta -1=1$ (see text). Again, the two measurements don't completely agree, but the slope from our data does seem to indicate such a behavior. \label{raw_aval_2d_fig}} \end{figure} Equation\ (\ref{2d_equ2}) is therefore one of the three scaling forms we use. The second scaling form is: \begin{equation} D_{int}(S,R)\ \sim\ S^{-(\tau + \sigma\beta\delta)}\ {\cal D}^{(int)(2)}_{-} \Bigl(S e^{-\tilde{a}/\sigma\nu|R_c-R|^2}\Bigr) \label{2d_equ3} \end{equation} shown in figure\ \ref{aval_2d_fig}c, with $\tau + \sigma\beta\delta=2.04$, $\tilde{a}/\sigma\nu=7.0$ (this implies that $\sigma\nu=0.49$), and $r=R_c-R$ with $R_c \equiv 0$ by assumption. And finally, the third scaling form we use is: \begin{equation} D_{int}(S,R)\ \sim\ S^{-(\tau + \sigma\beta\delta)}\ {\cal D}^{(int)(1)}_{-} \Bigl(S e^{-\tilde{b}/\sigma\nu|R_c-R|}\Bigr) \label{2d_equ6} \end{equation} shown in figure\ \ref{aval_2d_fig}d, with $\tau + \sigma\beta\delta=2.04$, $\tilde{b}/\sigma\nu = 4.0$ (which makes $\sigma\nu=0.51$), and $r=R_c-R$ with $R_c = 0.42$. Again, not only are all three collapses comparable, but the exponents extracted from them are as well. The exponent for the slope of the distribution is $\tau+\sigma\beta\delta=2.04$ for the three collapses, and the exponent combination $\sigma\nu$ is around $0.51$ (for the first collapse $\sigma=0.10$, while $\nu=5.25$ from the equivalent second moment collapse). Figures\ \ref{aval_2d_fig}(b-d) show that the scaling function ${\cal D}^{(int)}_{-}$ seems to be linear with slope one for small arguments (the grey lines have slope one) and that the constant term in the polynomial expansion is zero (or close to zero). This leads to a singular scaling function correction to the avalanche size distribution exponent $\tau+\sigma\beta\delta$ for small non--zero $\tilde X$: \begin{equation} D_{int}(S,R)\ \sim\ S^{-(\tau+\sigma\beta\delta)}\ {\cal D}_{-}^{(int)}(\tilde X)\ \sim\ S^{-(\tau+\sigma\beta\delta)+1} \label{2d_equ7} \end{equation} (Note that we could have used ${\cal D}^{(int)(1)}_{-}$ or ${\cal D}^{(int)(2)}_{-}$ as well.) Recall that because of the ``bump'' in the avalanche size distribution scaling function in $3$, $4$, and $5$ dimensions, and in mean field, the slope of the raw data curves did not agree with the value of the exponent $\tau+\sigma\beta\delta$. In $2$ dimensions, this is still true, but we also find a singular behavior for ${\cal D}_{-}^{(int)}(\tilde X)$, which changes the slope of the data curve for small $\tilde X$. In figure\ \ref{raw_aval_2d_fig}, an integrated avalanche size distribution curve for a system of $30000^2$ spins, at $R=0.65$, is plotted along with the linear fits to the data for small and large size $S$. For large $S$, the slope is close to but not equal to $2$, while for small $S$, the slope is close to one! The avalanche correlation data (see figure\ \ref{correl_2d_fig}a) is collapsed with three different scaling forms as well. These forms are analogous to the ones used for the second moments collapses, but with the distance $x$ taking the place of the system size $L$. The collapses and the extracted exponents from these three forms are again very similar, and only one of the collapses is shown in figure\ \ref{correl_2d_fig}b. The value of $\beta/\nu$ from these collapses is $0.03 \pm 0.06$. If we compare figure\ \ref{correl_2d_fig}b with the collapse of the avalanche correlation in $3$ dimensions (fig.\ \ref{correl_collapse_3d_fig}a), we find that the scaling function in $2$ dimensions seems to be singular with slope one for small distances, as is the integrated avalanche size distribution for small sizes. The spanning avalanches data are also analyzed using three scaling forms similar to those used for the second moments of the avalanche size distribution collapses. The exponent $\theta$ is poorly defined from these collapses (and is therefore not listed in Table\ \ref{conj_meas_2d_table}), although the data does collapse for the exact value: $\theta=0$. The three collapses for all the measurements we have done are very similar. This is not a surprise: it is always hard to distinguish large power laws ($\nu$ and $1/\sigma$ are large in the ``linear argument'' scaling form (eqns.\ (\ref{2d_equ01}) and (\ref{2d_equ2}))) from exponentials. Although some of the exponents have very different values in the three collapses, the average of the exponents from the three methods agree within the error bars with each method (see figure\ \ref{exp_compare}) and our conjectures. In conclusion, although we do not know the correct scaling form for the data in $2$ dimensions, the possible three scaling forms we mention give exponent values that are compatible with each other and with our conjectures (see Table\ \ref{conj_meas_2d_table}). (Table\ \ref{conj_2d_table} gives the conjectured values for the exponents that have not been measured in the collapses.) Much larger system sizes might be necessary to obtain more conclusive results. \begin{figure} \centerline{ \psfig{figure=Figures/Norm_Correl_d2_L30000_some_paper_THESIS.ps,width=3truein} } \nobreak \centerline{ \psfig{figure=Figures/Norm_Correl_d2_L30000_some_collapse_paper_THESIS.ps,width=3truein} } \nobreak \caption[Avalanche correlation curves in $2$ dimensions] {(a) {\bf Avalanche correlation function in $2$ dimensions,} integrated over the external field $H$, for several disorders $R$ and the system size $L=30000$. Only the curve with the smallest disorder is an average over several random field configuration. (b) Scaling collapse of the avalanche correlation curves in $2$ dimensions, for a system of $30000^2$ spins. The exponent values are: $\nu=5.25$ and $\beta/\nu=0$. The critical disorder is $R_c=0.54$, and $r=(R-R_c)/R$. Notice that for small $x|r|^\nu$, the scaling function looks singular with a power close to one (the straight line has a slope of one). \label{correl_2d_fig}} \end{figure} \section{Derivation of the various scaling forms and corrections} In this paper we make extensive use of scaling collapses. Many variations are important to us: Widom scaling, finite-size scaling, singular corrections to scaling, analytic corrections to scaling, rotating axes, and exponentially diverging correlation length scaling. The underlying theoretical framework for scaling is given by the renormalization group, developed by Wilson and Fisher\cite{WilsonFisher} in the context of equilibrium critical phenomena and by now well explicated in a variety of texts\cite{Goldenfeld,MaBinney,texts}. We have discovered that we can derive all the scaling forms and corrections that have been important to us from two simple hypotheses (found in critical regions): universality and invariance under reparameterizations. {\sl Universality} is the statement that two completely different systems will behave the same near their critical point\ \cite{note5} (for example, they can have exactly the same kinds of correlations). {\sl Reparameterization invariance} is the statement that smooth changes in the units or methods of measurement should not affect the critical properties. We use these properties to develop the scaling forms and corrections we use in this paper. Each example we cover will build on the previous ones while developing a new idea. For our first example, consider some property $F$ of a system at its critical point, as a function of a scale $x$. $F$ might be the spin-spin correlation function as a function of distance $x$ (or it might be the avalanche probability distribution as a function of size $x$, etc.) If two different experimental systems are at the same critical point, their $F$'s must agree. It would seem clear that they cannot be expected to be equal to one another: the overall scale of $F$ and the scale of $x$ will depend on the microscopic structure of the materials. The best one could imagine would be that \begin{equation} F_1(x_1) = A F_2( B x_2) \label{single} \end{equation} where $A$ would give the ratio of, say, the squared magnetic moment per domain of the two materials, and $B$ gives the ratio of the domain sizes. Now, consider comparing a system with itself, but with a different measuring apparatus. Universality in this self-referential sense would imply $F(x) = A F(B x)$, for suitable $A$ and $B$. If instead of using finite constant $A$ and $B$, we arrange for an infinitesimal change in the measurement of length scales, we find: \begin{equation} F(x) = (1-\alpha \epsilon)\ F\Bigl((1-\epsilon) x\Bigr) \label{small_single} \end{equation} where $\epsilon$ is small and $\alpha$ is some constant. Taking the derivative of both sides with respect to $\epsilon$ and evaluating it at $\epsilon = 0$, we find $-\alpha F = x F'$, so \begin{equation} F(x) \sim x^{-\alpha}. \label{power} \end{equation} The function $F$ is a power--law! The underlying reason why power--laws are seen at critical points is that power laws look the same at different scales. Now consider a new measurement with a distorted measuring apparatus. Now $F(x) \sim {\cal A}\Bigl[F\Bigl({\cal B}(x)\Bigr)\Bigr]$ where ${\cal A}$ and ${\cal B}$ are some nonlinear functions. For example, one might measure the number of microscopic domains $x$ flipped in an avalanche, or one might measure the total acoustic power ${\cal B}(x)$ emitted during the avalanche; these two ``sizes'' should roughly scale with one another, but nonlinear amplifications will occur while the spatial extent of the avalanche is small compared to the wavelength of sound emitted: we expand ${\cal B}(x) = B x + b_0 + b_1/x + \ldots$ Similarly, our microphone may be nonlinear at large sound amplitudes, or the absorption of sound in the medium may be nonlinear: ${\cal A}(F) = A F + a_2 F^2 + \ldots$ So, \begin{eqnarray} {\cal A}\Bigl[F\Bigl({\cal B}(x)\Bigr)\Bigr]\ \approx\ \nonumber \\ A \Bigl(F(B x) + F'(B x)(b_0 + b_1/x + \ldots)\ +\ \nonumber \\ F''(B x)(\ldots) + \ldots \Bigr) +\ a_2 F^2(B x)\ + \ldots \label{nonlinear_single} \end{eqnarray} We can certainly see that our assumption of universality cannot hold everywhere: for large $F$ or small $x$ the assumption of reparameterization invariance (\ref{nonlinear_single}) prevents any simple universal form. Where is universality possible? We can take the power-law form $F(x) \sim x^{-\alpha} = x^{\log A/\log B}$ which is the only form allowed by linear reparameterizations and plug it into (\ref{nonlinear_single}), and we see that all these nonlinear corrections are subdominant ({\it i.e.}, small) for large $x$ and small $F$ (presuming $\alpha>0$). If $\alpha>1$, the leading correction is due to $b_0$ and we expect $x^{-\alpha-1}$ corrections to the universal power law at small distances; if $0<\alpha<1$ the dominant correction is due to $a_2$, and we expect corrections of order $x^{-2\alpha}$. Thus our assumptions of universality and reparameterization invariance both lead us to the power-law scaling forms and inform us as to some expected deviations from these forms. Notice that the simple rescaling led to the universal power-law predictions, and that the more complicated nonlinear rescalings taught us about the dominant corrections: this will keep happening with our other examples. For our second example, let us consider a property $K$ of a system, as a function of some external parameter $R$, as we vary $R$ through the critical point $R_c$ for the material (so $r=R-R_c$ is small). $K$ might represent the second moment of the avalanche size distribution, where $R$ would represent the value of the randomness; alternatively $K$ might represent the fractional change in magnetization $\Delta M$ at the infinite avalanche $\ldots$ If two different experimental systems are both near their critical points ($r_1$ and $r_2$ both small), then universality demands that the dependence of $K_1$ and $K_2$ on ``temperature'' $R$ must agree, up to overall changes in scale. Thus, using a simple linear rescaling $K(r) = (1-\mu \epsilon) K\Bigl((1-\epsilon) r\Bigr)$ leads as above to the prediction \begin{equation} K(r) = r^{-\mu}. \label{linear_single} \end{equation} Now let us consider nonlinear rescalings, somewhat different than the one discussed above. In particular, the nonlinearity of our measurement of $K$ can be dependent on $r$. So, \begin{equation} {\cal A}_r\Bigl(K(r)\Bigr) = a_0 + a_1 r + a_2 r^2 + \ldots + a_{01} K(r) + \ldots \label{nonlinear_k} \end{equation} If $\mu>0$, these analytic corrections don't change the dominant power law near $r=0$. However, if $\mu<0$, all the terms $a_n$ for $n<-\mu$ will be more important than the singular term! Only after fitting them to the data and subtracting them will the residual singularity be measurable. For the fractional change in magnetization: $\Delta M \sim r^\beta$ has $0< \beta < 1$ (at least above three dimensions), so we might think we need to subtract off a constant term $a_0$, but $\Delta M = 0$ for $R \ge R_c$, so $a_0$ is zero. On the other hand, in a previous paper\cite{Dahmen1}, we discussed the singularity in the area of the hysteresis loop: $Area \sim r^{2-\alpha}$, where $2-\alpha = \beta + \beta \delta$ is an analogue to the specific heat in thermal systems. Since $\alpha$ is near zero (slightly positive from our estimates of $\beta$ and $\delta$ in 3, 4, and 5 dimensions), measuring it would necessitate our fitting and subtracting three terms (constant, linear, and quadratic in $r$): we did not measure the area for that reason. For our third example, let's consider a function $F(x,r)$, depending on both a scale $x$ and an external parameter $r$. For example, $F$ might be the probability $D_{int}$ that an avalanche of size $x$ will occur during a hysteresis loop at disorder $r=R-R_c$. Universality implies that two different systems must have the same $F$ up to changes in scale, and therefore that $F$ measured at one $r$ must have the same form as if measured at a different $r$. To start with, we consider a simple linear rescaling: \begin{equation} F(x,r) = (1 - \alpha \epsilon)\ F\Bigl( (1-\epsilon) x, (1+\zeta \epsilon) r \Bigr). \label{linear_double} \end{equation} Taking the derivative of both sides with respect to $\epsilon$ gives a partial differential equation that can be manipulated to show $F$ has a scaling form. Instead, we change variables to a new variable $y = x^\zeta r$ (which satisfies $y'=y$ to order $\epsilon$). If $\tilde F(x,y) \equiv F(x,r)$ is our function measured in the new variables, then \begin{equation} F(x,r) = \tilde F(x,y) = (1-\alpha \epsilon)\ \tilde F\Bigl((1-\epsilon)x,y\Bigr) \label{linear_double_tilde} \end{equation} and $-\alpha \tilde F = x \, \partial \tilde F/\partial x$ shows that at fixed $y$, $F\sim x^{-\alpha}$, with a coefficient ${\cal F}(y)$ which can depend on $y$. Hence we get the scaling form \begin{equation} F(x,r) \sim x^{-\alpha}\ {\cal F}(x^\zeta r). \label{double_scaling} \end{equation} This is just Widom scaling. The critical exponents $\alpha$ and $\zeta$, and the scaling function ${\cal F}(x^\zeta r)$ are universal (two different systems near their critical point will have the {\sl same} critical exponents and scaling functions). We don't need to discuss corrections to scaling for this case, as they are similar to those discussed above and below (and because none were dominant in our cases). Notice that if we sit at the critical point $r=0$, the above result reduces to equation (\ref{power}) so long as ${\cal F}(0)$ is not zero or infinity. If, on the other hand, ${\cal F}(y) \sim y^n$ as $y \to 0$, the two-variable scaling function gives a singular correction to the power--law near the critical point: $F(x,r) \sim x^{-\alpha}\ {\cal F}(x^\zeta r) \sim x^{-\alpha + n \zeta}\ $ for $x <\!< r^{-1/\zeta}$: only when $x \sim r^{-1/\zeta}$ will the power-law $x^{-\alpha}$ be observed. This is what happened in two dimensions to the integrated avalanche size distribution (figures\ \ref{aval_2d_fig} and \ref{raw_aval_2d_fig}) and the avalanche correlation functions (figure\ \ref{correl_2d_fig}b). For the fourth example, we address finite-size scaling of a property $K$ of the system, as we vary a parameter $r$. If we measure $K(r,L)$ for a variety of sizes $L$ (say, all with periodic boundary conditions), we expect (in complete analogy to (\ref{double_scaling})) \begin{equation} K(r,L) \sim r^{-\mu}\ {\cal K}(r L^{1/\nu}). \label{finite_size_scaling} \end{equation} Now, suppose our ``thermometer'' measuring $r$ is weakly size-dependent, so the measured variable is ${\cal C}(r) = r + c/L + c_2/L^2 + \ldots$\ The effects on the scaling function is \begin{eqnarray} K\Bigl({\cal C}(r),L\Bigr) \sim r^{-\mu}\ \times \nonumber \\ \Bigl({\cal K}(r L^{1/\nu})\ +\ \nonumber \\ (c L^{1/\nu-1} + c_2 L^{1/\nu-2})\ {\cal K}'(r L^{1/\nu}) + \ldots \Bigr). \label{finite_size_corrections_to_scaling} \end{eqnarray} In two and three dimensions, $\nu>1$ and these correction terms are subdominant. In four and five dimensions, we find $1/2 < \nu < 1$, so we should include the term multiplied by $c$ in equation (\ref{finite_size_corrections_to_scaling}). However, we believe this first term is zero for our problem. For a fixed boundary problem (all spins ``up'' at the boundary) with a first order transition, there is indeed a term like $c/L$ in $r(L)$\ \cite{finite_size_first_order}. At a critical transition, the leading correction to $r(L)$ can be $c/L$ or a higher power of $L$ ($1/L^2$ and so on). This seems to depend on the model studied, the geometry of the system, and the boundary conditions (free, periodic, ferromagnetic, $\ldots$)\ \cite{finite_size_critical}. Furthermore, for the same kind of model, the coefficient $c$ itself depends on the geometry and boundary conditions, and it can even vanish, which leaves only higher order corrections. In a periodic boundary conditions problem like ours, we expect that the correction is smaller than $c/L$. Our finite-size scaling collapses for spanning avalanches $N$, the second moments $\langle S^2\rangle$, and the magnetization jump $\Delta M$, were successfully done by letting $c=0$. For the fifth example, consider a property $K$ depending on two external parameters: $r$ (the disorder for example) and $h$ (could be the external magnetic field $H-H_c$). Analogous to (\ref{double_scaling}), $K$ should then scale as \begin{equation} K(r,h) \sim r^{-\mu}\ {\cal K}(h/r^{\beta\delta}). \label{double_scaling_2} \end{equation} Consider now the likely dependence of the field $h$ on the disorder $r$. A typical system will have a measured field which depends on the randomness: $\tilde{\cal C}(h) = h + b\, r + b_2 r^2 + \ldots$ (Corresponding nonlinearities in the effective value of $r$ are subdominant.) This system will have \begin{eqnarray} K\Bigl(r,\tilde{\cal C}(h)\Bigr)\ =\ r^{-\mu}\ \times \nonumber \\ \Bigl( {\cal K}(h/r^{\beta\delta}) + (b\, r + b_2 r^2)\ r^{-\beta\delta}\ {\cal K}'(h/r^{\beta\delta}) \Bigr). \label{rotated_double_scaling} \end{eqnarray} Now, for our system $1 < \beta \delta < 2$ for dimensions three and above. This means that the term multiplied by $b$ is dominant over the critical scaling singularity: unless one shifts the measured $h$ to the appropriate $h'=h + b\,r$, the curves will not collapse ({\it e.g.}, the peaks will not line up horizontally). We measure this (non-universal) constant for our system (Table\ \ref{RH_table}), using the derivative of the magnetization with field $dM/dH(r,h)$. The magnetization $M(r,h)$ and the correlation length $\xi(r,h)$ should also collapse according to equation (\ref{double_scaling_2}) (but with $h + b\,r$ instead of $h$); we don't directly measure the correlation length, and the collapse of $M(r,h)$ in figure\ \ref{3d_MofH_fig}b includes the effects of the tilt $b$. In two dimensions, $\beta \delta$ is large (probably infinite), so in principle we should need an infinite number of correction terms: in practise, we tried lining up the peaks in the curves (with no correction terms); because we did not know $\beta$ (which we usually obtained from $\Delta M$, which gives $\beta/\nu=0$ in two dimensions), we failed to extract reliable exponents in two dimensions from $dM/dH$. For the sixth example, suppose $F$ depends on $r$, $h$, and a size $x$. Then from the previous analysis, we expect \begin{equation} F(x,r,h) \sim x^{-\alpha}\ {\cal F}(x^\zeta r,\, h/r^{\beta \delta}). \label{triple_scaling} \end{equation} Notice that universality only removes one variable from the scaling form. One could in practice do two--variable scaling collapses (and we believe someone has probably done it), but for our purposes these more general scaling forms are used by fixing one of the variables. For example, we measure the avalanche size distribution at various values of $h$ (binned in small ranges), at the critical disorder $r=0$. We can make sense of equation (\ref{triple_scaling}) by changing variables from $h/r^{\beta \delta}$ to $x^{\zeta \beta \delta} h$: \begin{equation} F(x,r,h) \sim x^{-\alpha} \tilde{\cal F}(x^\zeta r,\, x^{\zeta \beta \delta} h). \label{triple_scaling_nice} \end{equation} Before we can set $r=0$, we must see what are the possible corrections to scaling in this case. If the disorder $r$ depends on the field, then instead of the variable $r$, we must use $r + a h$ (the analysis is analogous to the one in example five; other corrections are subdominant). Setting $r=0$ now, leaves $F$ dependent on its first variable, as well as the second: \begin{eqnarray} F(x,r,h) & \sim\ x^{-\alpha}\ \tilde{\cal F}(x^\zeta (a h),\, x^{\zeta \beta \delta} h) \approx\ x^{-\alpha}\ \times \nonumber \\ & \Bigl(\tilde{\cal F}(0,\ x^{\zeta \beta \delta} h) + \nonumber \\ & a h x^\zeta\ \tilde{\cal F}^{(1,0)}(0,\, x^{\zeta \beta \delta} h) \Bigr), \label{triple_scaling_reduced_corrections} \end{eqnarray} where $\tilde{\cal F}^{(1,0)}$ is the derivative of $\tilde{\cal F}$ with respect to the first variable (keeping the second fixed). For the binned avalanche size distribution, $x^\zeta$ is $S^\sigma$, where $0 \le \sigma < 1/2$ as we move from two dimensions to five and above. Thus, the correction term will only be important for rather large avalanches, $S > h^{-1/\sigma}$, so long as we are close to the critical point. Expressed in terms of the scaling variable, important corrections to scaling occur if the scaling variable $X = S^{\sigma\beta\delta} h > h^{1-\beta\delta}$. For us, $\beta \delta > 3/2$, and we only use fields near the critical field ($h < 0.08$), so the corrections will become of order one when $X=4$ for the largest $h$ we use. In $3$ and $4$ dimensions, this correction does not affect our scaling collapses, while in $5$ dimensions some of the data needs this correction. We have tried to avoid this problem (since we don't measure our data such that it can be used in a two--variable scaling collapse) by concentrating on collapsing the regions in our data curves where this correction is negligible. A similar analysis can be done for the avalanche time distribution, which has two ``sizes'' $S$ and $t$ and one parameter $r$ which is set to zero; because we integrate over the field $h$ the correction in (\ref{triple_scaling_reduced_corrections}) does not occur, and other scaling corrections are small. Finally, we discuss the unusual exponential scaling forms we developed to collapse our data in two dimensions. If we assume that the critical disorder $R_c$ is zero {\it and} that the linear term in the rescaling of $r$ vanishes ($\zeta \epsilon r$ in equation\ (\ref{linear_double}) vanishes), then from symmetry the correction has to be cubic, and equation\ (\ref{linear_double}) becomes: \begin{equation} F(x,r) = (1 - \alpha \epsilon)\ F\Bigl( (1-\epsilon) x,\, (1+ k \epsilon\, r^2) r \Bigr). \label{exponential_scaling_cubic} \end{equation} with $k$ (which is not universal) and $\alpha$ constants, and $\epsilon$ small. Taking the derivative of both sides with respect to $\epsilon$ and setting it equal to zero gives a partial differential equation for the function $F$. To solve for $F$, we do a change of variable: $(x,r) \rightarrow (x,y)$ with $y=x\ e^{-a^*/r^2}$. The constant $a^*$ is determined by requiring that $y$ rescales onto itself to order $\epsilon$: we find $a^*=1/2\,k$. We then have: \begin{equation} 0 = -\alpha\ \tilde{F}(x,y) - {\partial{\tilde{F}} \over \partial{x}}\ x \label{exponential_scaling_cubic_2} \end{equation} which gives \begin{equation} F(x,r) = x^{-\alpha}\ \tilde{\cal{F}}\Bigl(xe^{-1/2\,k\,r^2}\Bigr). \label{exponential_scaling_cubic_3} \end{equation} This is one of the forms we use in $2$ dimensions for the scaling collapse of the second moments $\langle S^2 \rangle_{int}$, the avalanche size distribution $D_{int}$ integrated over the field $H$, the avalanche correlation $G_{int}$, and the spanning avalanches $N$. We use another form too which is obtained by assuming that the critical disorder $R_c$ is not zero but that the linear term in the rescaling of $r$ still vanishes. Instead of equation\ (\ref{exponential_scaling_cubic}), we have: \begin{equation} F(x,r) = (1 - \alpha \epsilon)\ F\Bigl( (1-\epsilon) x,\, (1+ l \epsilon\, r) r \Bigr). \label{exponential_scaling_square} \end{equation} The function $F$ becomes: \begin{equation} F(x,r) = x^{-\alpha}\ \tilde{\cal{F}}\Bigl(xe^{-1/l\,r}\Bigr). \label{exponential_scaling_square_2} \end{equation} The corrections to scaling for the last two forms (equations\ (\ref{exponential_scaling_cubic_3}) and (\ref{exponential_scaling_square_2})) are similar to the ones discussed above. They are all are subdominant. \section{Full derivation of the mean field scaling form for the integrated avalanche size distribution} The mean field scaling form for the integrated avalanche size distribution $D_{int}(S,R)$ was obtained in section IV A using the {\it scaling form} of the avalanche size distribution $D(S,R,H)$. The scaling form for $D_{int}(S,R)$ can also be obtained by integrating the avalanche probability distribution $D(S,t)$ (derived originally in\cite{Sethna}) directly: \begin{equation} D_{int} (S,R) = \int_{-\infty}^{+\infty} \rho(-JM-H)\ D(S,t)\ dH \label{apA1_eq1} \end{equation} where $\rho (-JM-H)$ is the probability distribution for the random fields, and $\rho (-JM-H)\ dH$ is the probability for a spin to flip between fields $-JM(H) - H$ and $-JM(H+dH) - (H+dH)$. $D(S,t)$ is the probability of having an avalanche of size $S$, a small ``distance'' $t \equiv 2J \rho (-JM-H) - 1$ from the infinite avalanche at $\rho (-JM-H) = 1/2J$, given that a spin has flipped at $-JM - H$\cite{Sethna,Dahmen1}. (The scaling form for the non-integrated avalanche size distribution $D(S,R,H)$ (eqn.\ref{int_aval0}) is obtained from $D(S,t)$ by expressing $t$ as a function of $R$ and $H$ \cite{Sethna,Dahmen1}). $J$ is the coupling of a spin to all others in the system, $H$ is the external magnetic field, and $R$ is the disorder. The advantage of this procedure is that we can find out something about the scaling function $\bar {\cal D}_{-}^{(int)}$. The average mean field magnetization $M$ and the avalanche probability distribution $D(S,t)$ are given by \cite{Sethna,Dahmen1}: \begin{equation} M(H,R) = 1 - 2 \int_{-\infty}^{-JM(H)-H} \rho(f)\ df, \label{apA1_eq2} \end{equation} and \begin{equation} D(S,t) = {S^{S-2} \over (S-1)!}\ (t+1)^{S-1}\ e^{-S(t+1)} \label{apA1_eq3} \end{equation} To solve equation (\ref{apA1_eq1}), let's define the variable $y=(-JM-H)/({\sqrt 2}\ R)$ and rewrite the integral as: \begin{eqnarray} D_{int} (S,R)\ =\ {\sqrt 2}\ R\ \times \nonumber \\ \Biggl[\int_{-\infty}^{+\infty} \rho ({\sqrt 2}Ry)\ D\Bigl(S,\ 2J \rho ({\sqrt 2}Ry) - 1\Bigr) \times \nonumber \\ \ \Bigl(1 - 2J \rho ({\sqrt 2}Ry)\Bigr)\ dy \Biggr], \label{apA1_eq4} \end{eqnarray} where we have used: \begin{equation} {dy \over dH} = {1 \over {\sqrt 2}\ R}\ \biggl(-J\ {2\ \rho(-JM-H) \over {1-2J \rho(-JM-H)}} -1\biggr) \label{apA1_eq5} \end{equation} Since we are interested in the behavior of the integrated avalanche distribution for large sizes, the factorial in equation (\ref{apA1_eq3}) can be expanded using Stirling's formula. To first order, we have: \begin{equation} (S-1)!\ \approx\ {S^S\ {\sqrt {2 \pi}} \over e^S\ {\sqrt S}} \label{apA1_eq6} \end{equation} Substituting this and the random field distribution function $\rho$, \begin{equation} \rho ({\sqrt 2}Ry) = {1 \over {\sqrt {2 \pi}} R}\ e^{-y^2}, \label{apA1_eq7} \end{equation} in equation (\ref{apA1_eq4}), we obtain: \begin{eqnarray} D_{int}(S,R)\ \approx\ C\ \biggl({R_c \over R}\biggr)^S\ \times \nonumber \\ \int_{-\infty}^{+\infty} e^{-S\ \bigl(y^2 + {R_c \over R}\ e^{-y^2}\bigr)} \ \ \biggl(1-{R_c \over R}\ e^{-y^2}\biggr)\ dy \label{apA1_eq8} \end{eqnarray} where $C=S^{-{3 \over 2}}\ e^S\ R_c/(\pi R {\sqrt 2})$, and $S$ is large. For disorders above but close to the critical disorder $R_c$, we have: \begin{eqnarray} \biggl({R_c \over R}\biggr)^S\ =\ e^{S\ log\bigl({R_c \over R}\bigr)}\ \approx\ \nonumber \\ e^{S\ \Bigl(-\ {\bigl(1-{R_c \over R}\bigr) \over 1} - \ {\bigl(1-{R_c \over R}\bigr)^2 \over 2} - {\bigl(1 - {R_c \over R}\bigr)^3 \over 3} - ... \Bigr)} \label{apA1_eq9} \end{eqnarray} If we assume that only terms up to $S\ (1-R_c/R)^2$ are important (terms like $S\ (1-R_c/R)^3$ and $S\ (1-R_c/R)^4$ go to zero as $R \rightarrow R_c$), and we note that the integrand in equation (\ref{apA1_eq8}) is an even function of $y$, equation (\ref{apA1_eq8}) becomes: \begin{eqnarray} D_{int}(S,R)\ \approx\ 2\ C\ \times \nonumber \\ \Biggl[ \int_{0}^{+\infty} e^{-S\ \Bigl({\bigl(1-{R_c \over R}\bigr) \over 1} + {\bigl(1-{R_c \over R}\bigr)^2 \over 2} +y^2 + {R_c \over R}\ e^{-y^2}\Bigr)} \times \nonumber \\ \ \ \biggl(1-{R_c \over R}\ e^{-y^2}\biggr)\ dy \Biggr] \label{apA1_eq10} \end{eqnarray} The asymptotic behavior of the above integral, as $S \rightarrow \infty$, is obtained using Laplace's method\cite{BenderOrszag}. The idea is as follows. The asymptotic behavior as $S \rightarrow \infty$ of the integral: \begin{equation} I(S) = \int_{a}^{b} f(x)\ e^{S \phi (x)}\ dx \label{apA1_eq11} \end{equation} can be found by integrating over a small region $[c-\epsilon,\ c+\epsilon]$ (instead of the interval $[a,b]$) around the maximum of the function $\phi (x)$ at $x=c$, since in the asymptotic expansion, the largest contribution to the integral will be from this region. The corrections will be exponentially small. The maximum of $\phi$ must be in the interval $[a,b]$, $f(x)$ and $\phi(x)$ are assumed to be real continuous functions, and $f(c) \ne 0$. $f(x)$ and $\phi (x)$ can now both be expanded around $x=c$, and the integral solved. Often the integral is easier to handle if the limit of integration is extended to infinity. This will add only exponentially small corrections in the asymptotic limit of $S \rightarrow \infty$. Let's apply this method to equation (\ref{apA1_eq10}). The function in the exponential has a maximum at $y=0$. The function $\biggl(1-{R_c \over R}\ e^{-y^2}\biggr)$ is not zero there if $R \ne R_c$. We can thus expand both functions in the integral of equation (\ref{apA1_eq10}) around $y=0$. Defining $u=y^2{\sqrt S}$, we obtain: \begin{eqnarray} D_{int}(S,R)\ \approx\ C_{1}\ \times \nonumber \\ \Biggl[ \int_{0}^{\epsilon} e^{-{\sqrt S}\ \Bigl(\bigl(1-{R_c \over R}\bigr) u\ +\ {u^2 \over {2 {\sqrt S}}}{R_c \over R} -\ {u^3 \over {6\ S^2}}{R_c \over R} + ...\Bigr)}\ \times \nonumber \\ \biggl(\Bigl(1-{R_c \over R}\Bigr) + {R_c \over R} {u \over {\sqrt S}} -\ {R_c \over 2R} {u^2 \over S} + ...\biggr) \ {du \over {\sqrt u}} \Biggr] \label{apA1_eq12} \end{eqnarray} where \begin{equation} C_{1} = {1 \over \pi {\sqrt 2}}\ {R_c \over R}\ S^{-{9 \over 4}} \ \ e^{-{S \over 2}\ \bigl(1-{R_c \over R}\bigr)^2}, \label{apA1_eq13} \end{equation} $S$ is large, $R$ is close to but not equal to $R_c$, and only terms up to $S (1-R_c/R)^2$ are non-vanishing. In the asymptotic limit of $S \rightarrow \infty$ we can ignore terms with powers of $S$ in the denominator, and look at the distribution for $R$ close to $R_c$. To first order in $r = (R_c-R)/R$, $R_c/R \approx 1$ and $1-R_c/R \approx -r$, which gives: \begin{eqnarray} D_{int}(S,R)\ \approx\ {1 \over \pi {\sqrt 2}}\ S^{-{9 \over 4}} \ e^{-{S \over 2}\ (-r)^2}\ \times \nonumber \\ \int_{0}^{\infty} e^{\Bigl(-(-r){\sqrt S}\ u\ -\ {u^2 \over 2}\Bigr)}\ \ \Bigl(-r {\sqrt S} + u\Bigr) \ {du \over {\sqrt u}} \label{apA1_eq14} \end{eqnarray} where we have expanded the integration to infinity. As mentioned above, this will only add exponentially small corrections in the asymptotic limit of $S \rightarrow \infty$. Equation\ (\ref{apA1_eq14}) is the integrated avalanche size distribution in mean field for large sizes $S$, and finite $S r^2$. We see right away that it gives the correct scaling form: \begin{equation} D_{int}(S,R)\ \sim\ S^{-{9 \over 4}}\ \bar{\cal D}_{\pm}^{(int)} \Bigl({\sqrt S}\ |r|\Bigr) \label{apA1_eq15} \end{equation} where $\pm$ indicates the sign of $r$, the exponent $\tau+\sigma\beta\delta$ and $\sigma$ are $9/4$ and $1/2$ respectively, and the scaling function $\bar{\cal D}_{\pm}^{(int)}$ is: \begin{equation} \bar{\cal D}_{\pm}^{(int)} \Bigl({\sqrt S}\ |r|\Bigr) = e^{-{({\sqrt S}\ |r|)^2 \over 2}}\ \bar{\cal F}_{\pm} \Bigl({\sqrt S}\ |r|\Bigr). \label{apA1_eq16} \end{equation} The function $\bar{\cal F}_{\pm} \Bigl({\sqrt S}\ |r|\Bigr)$ is proportional to the integral in equation\ (\ref{apA1_eq14}). Note that the above result is equivalent to the one obtained (eqn.\ \ref{int_aval3}) by integrating the {\it scaling form} of $D(S,R,H)$ over the field $H$. What is the behavior of the scaling function $\bar{\cal D}_{-}^{(int)}(X)$ for small and large positive arguments $X={\sqrt S}\ (-r) > 0$ ($R > R_c$)? From equations\ (\ref{apA1_eq14}) and (\ref{apA1_eq16}), for small arguments we have a polynomial in $X$: \begin{equation} \bar{\cal D}_{-}^{(int)}(X) \approx A+BX+CX^2+{\cal O}(X^3) \label{apA1_eq17} \end{equation} These parameters can be calculated numerically. We obtain in mean field: \begin{eqnarray} {\bar{\cal D}_{-}^{(int)}}\ & \approx\ & 0.232+0.243 X-0.174 X^2- \nonumber \\ & & 0.101 X^3+0.051 X^4 \label{dist_mf2} \end{eqnarray} On the other hand, for large arguments we find: \begin{equation} \bar{\cal D}_{-}^{(int)}(X) \approx {\pi}^{1/2} e^{-{X^2 \over 2}}\ {\sqrt X} \Bigl( 1 + {\cal O} (X^{-2})\Bigr) \label{apA1_eq18} \end{equation} In general, for all dimensions, in equation\ (\ref{apA1_eq18}) the exponential is of $X^{1/\sigma}$ ($1/\sigma = 2$ in mean field), since the exponent $\sigma$ gives the exponential cutoff to the power law distribution for large $X$, and the power of $X$ is $\beta$ ($\beta = 1/2$ in mean field). One can see the latter by expanding the distribution function $D_{int}(S,R)$ in terms of $1/S$ ($S$ is large), analogous to\ \cite{Griffiths}: \begin{equation} D_{int}(S,R) = \sum_{n=1}^{\infty} f_n (r)\ S^{-n} \label{apA1_equ18a} \end{equation} Since $X=$ $S^\sigma (-r)$, then we can write $S=$ $X^{1/\sigma}\ (-r)^{-1/\sigma}$ and obtain: \begin{equation} D_{int}(S,R) = \sum_{n=1}^{\infty} f_n (r)\ X^{-n/\sigma} (-r)^{n/\sigma} \label{apA1_equ18b} \end{equation} The scaling function $\bar{\cal D}_{-}^{(int)}(X)$ scales like $S^{(\tau+\sigma\beta\delta)}\ \times D_{int}(S,R)$: \begin{eqnarray} \bar{\cal D}_{-}^{(int)}(X) \sim\ \Biggl[ \sum_{n=1}^{\infty} f_n (r)\ X^{-n/\sigma} X^{(\tau+\sigma\beta\delta)/\sigma}\ \times \nonumber \\ (-r)^{n/\sigma} (-r)^{-(\tau+\sigma\beta\delta)/\sigma} \Biggr] \label{apA1_equ18c} \end{eqnarray} and since it is only a function of $X$, it must satisfy: \begin{equation} \bar{\cal D}_{-}^{(int)}(X) \sim\ \sum_{n=1}^{\infty} g_n\ X^{-n/\sigma} X^{(\tau+\sigma\beta\delta)/\sigma} \label{apA1_equ18d} \end{equation} where $g_n$ is independent of $r$. The exponent combination $(\tau+\sigma\beta\delta)/\sigma$ can be rewritten as: \begin{equation} {\tau+\sigma\beta\delta \over \sigma} = {2 \over \sigma} + {\tau+\sigma\beta\delta - 2 \over \sigma} = {2 \over \sigma} + \beta \label{apA1_equ18e} \end{equation} where we have used the scaling relation\cite{Dahmen1,Dahmen3}: $\beta - \beta\delta = (\tau - 2)/\sigma$. Thus we have for the scaling function $\bar{\cal D}_{-}^{(int)}(X)$: \begin{equation} \bar{\cal D}_{-}^{(int)}(X) \sim\ X^\beta \sum_{n=-1}^{\infty} g_n\ X^{-n/\sigma} = X^\beta\ {\cal K}(X^{1/\sigma}) \label{apA1_equ18f} \end{equation} which shows (compare to equation\ (\ref{apA1_eq18})) that the power of $X$ is indeed the exponent $\beta$. We have used the results of the expansion of the mean field scaling function ${\bar {\cal D}}_{-}^{(int)} (X)$ for small and large parameters (equations\ (\ref{apA1_eq17}) and (\ref{apA1_eq18})), to build a fitting function to the integrated avalanche size distribution scaling functions in $2$, $3$, $4$, and $5$ dimensions, described in section IV~ B. Finally, note from equation\ (\ref{apA1_eq17}) that the scaling function: \begin{equation} {\cal D}_{-}^{(int)} (S r^2) = e^{-{(S r^2) \over 2}}\ {\cal F}_{-}(S r^2) \label{apA1_eq19} \end{equation} used earlier in reference\cite{Dahmen1}, is not analytic for small arguments $S r^2$, from which we conclude that the appropriate scaling variable should be ${\sqrt S}\ (-r)$ and not $S r^2$. (Notice that this no longer seems true in two dimensions; see section on $2$ dimensional results.) \section{Derivation of the mean field scaling form for the spanning avalanches} We have defined earlier a mean field spanning avalanche to be an avalanche larger than $\sqrt {S_{mf}}$, where $S_{mf}$ is the {\it total} size of the system. We want to derive the scaling form for the number of such avalanches in half of the hysteresis loop (for $H$ from $-\infty$ to $+\infty$) as a function of the system size $S_{mf}$ and the disorder $R$. The number of mean field spanning avalanches is proportional to the probability of having avalanches of size larger than $\sqrt {S_{mf}}$. Since we want the number of spanning avalanches, we need to multiply this probability by the total number of avalanches. For large system sizes, this scales with the system size $S_{mf}$ (corrections are subdominant). We thus obtain by integrating over equation\ (\ref{apA1_eq15}) (which gives the scaling form for the probability distribution of avalanches of size $S$ in the hysteresis loop): \begin{eqnarray} N_{mf}(S_{mf},R)\ \sim\ S_{mf}\ \times \nonumber \\ \int_{\sqrt {S_{mf}}}^{\infty}\ S^{-{9 \over 4}}\ e^{-{\bigl({\sqrt S} |r|\bigr)^2 \over 2}}\ \bar{\cal F}_{\pm}\Bigl({\sqrt S} |r|\Bigr)\ dS \label{apB_eq1} \end{eqnarray} Let's define $u={\sqrt S} |r|$, then equation\ (\ref{apB_eq1}) can be written as: \begin{eqnarray} N_{mf}(S_{mf},R)\ \sim\ 2\ S_{mf}\ |r|^{5 \over 2} \times \nonumber \\ \int_{|r| S_{mf}^{1 \over 4}}^{\infty}\ u^{-{7 \over 2}}\ e^{-{u^2 \over 2}}\ \bar{\cal F}_{\pm}(u)\ du \label{apB_eq2} \end{eqnarray} The integral ${\cal I}$ is a function of $S_{mf}^{1 \over 4} |r|$ only, and we can write it as: \begin{equation} {\cal I} = \Bigl(S_{mf}^{1 \over 4} |r|\Bigr)^{-{5 \over 2}}\ {\cal N}_{\pm}^{mf}\Bigl(S_{mf}^{1 \over 4} |r|\Bigr) \label{apB_eq3} \end{equation} to obtain the scaling form for the number $N_{mf}$ of mean field spanning avalanches: \begin{equation} N_{mf}(S_{mf},R)\ \sim\ S_{mf}^{3 \over 8}\ {\cal N}_{\pm}^{mf}\Bigl(S_{mf}^{1 \over 4} |r|\Bigr) \label{apB_eq4} \end{equation} From this scaling form, we can extract the exponents $\tilde \theta =3/8$, and $1/\tilde \nu=1/4$. This form is used for collapses of the spanning avalanche curves in mean field (see mean field section). \subsubsection{Avalanche Size Distribution} \paragraph{Integrated Avalanche Size Distribution} In our model the spins often flip in avalanches, which are collective flips of neighboring spins at a constant external field $H$. These avalanches come in different sizes. The integrated avalanche size distribution is the size distribution of all the avalanches that occur in one branch of the hysteresis loop (for $H$ from $-\infty$ to $\infty$). Figure\ \ref{aval_3fig}\cite{Perkovic} shows some of the raw data (thick lines) in $3$ dimensions. Note that the curves follow a power law behavior over several decades. Even $50\%$ away from criticality (at $R=3.2$), there are still two decades of scaling, which implies that the critical region is large. In experiments, a few decades of scaling could be interpreted in terms of self-organized criticality (SOC). However, our model and simulation suggest that several decades of power law scaling can still be present rather ${\it far}$ from the critical point (note that the size of the critical region is non--universal). In the figure, the cutoff in the power law which diverges as the critical disorder $R_c$ is approached ($R_c=2.16$ in $3$ dimensions), is a signature that the system is away from criticality, and that a parameter can be tuned (here $R$) to bring it to the transition. This cutoff scales as $S \sim |r|^{-1/\sigma}$, where $S$ is the avalanche size and $r=(R_c-R)/R$ is the reduced disorder. The power law for the curves of figure\ \ref{aval_3fig} can be obtained through scaling collapses. A plot is shown in the inset of figure\ \ref{aval_3fig}. The scaling form is (see mean field section) \begin{equation} D_{int}(S,R) \sim S^{-(\tau+\sigma\beta\delta)}\ \bar{{\cal D}}^{(int)}_{-} (S^{\sigma}|r|) \label{aval_3d_eq} \end{equation} where $\bar{{\cal D}}_{-}^{(int)}$ is the scaling function (the $-$ sign indicates that the collapsed curves are for $R > R_c$). The critical exponents $\tau+\sigma\beta\delta=2.03$ and $\sigma=0.24$ are obtained from collapses and linear extrapolation of the extracted values to $R=R_c$ (figures\ \ref{3d_aval_expfig}a and \ref{3d_aval_expfig}b), as was done in mean field. (Although the ``real'' scaling variables are $r^{\prime}$ and $h^{\prime}$, when integrating over the field $H$ we recover the same form as in mean field; see appendix A.) Table\ \ref{measured_exp_table} lists all the exponents extracted from scaling collapses, and extrapolated to $R \rightarrow R_c$ and $1/L \rightarrow 0$. \begin{figure} \centerline{ \psfig{figure=Figures/Aval_histo_d3_L320_paper_norm_THESIS.ps,width=3truein} } \caption[Integrated avalanche size distribution curves in $3$ dimensions] {{\bf Avalanche size distribution integrated over the field $H$ in $3$ dimensions}, for $320^3$ spins and disorders $R=4.0, 3.2$, and $2.6$. The last curve is at $R=2.25$, for a $1000^3$ spin system. The $320^3$ curves are averages over up to $16$ initial random field configurations. All curves are smoothed by $10$ data points before they are collapsed. The inset shows the scaling collapse of the integrated avalanche size distribution curves in $3$ dimensions, using $r=(R_c-R)/R$, $\tau+\sigma\beta\delta=2.03$, and $\sigma=0.24$, for sizes $160^3$, $320^3$, $800^3$, and $1000^3$, and disorders ranging from $R=2.25$ to $R=3.2$ ($R_c = 2.16$). The two top curves in the collapse, at $R=3.2$, show noticeable corrections to scaling. The thick dark curve through the collapse is the fit to the data (see text). In the main figure, the distribution curves obtained from the fit to the collapsed data are plotted (thin lines) alongside the raw data (thick lines). The straight dashed line is the expected asymptotic power law behavior: $S^{-2.03}$, which does not agree with the measured slope of the raw data due to the shape of the scaling function (see text). \label{aval_3fig}} \end{figure} We have mentioned earlier that the mean field scaling function $\bar{\cal D}_{-}^{(int)}(X)$, with $X=S^{\sigma}|r|$ and $r<0$, is a polynomial for small $X$ and gives an exponential in $X^{1/\sigma}$ ($1/\sigma=2$ in mean field) multiplied by $X^\beta$ ($\beta=1/2$ in mean field) for large $X$ (see mean field section and appendix B). As we have done in mean field, we can try to fit the scaling function $\bar{\cal D}_{-}^{(int)}$ in dimensions $5$ and below with a product of a polynomial and an exponential function. This is done in $3$ dimensions in the inset of figure\ \ref{aval_3fig} (thick black line through the data). The {\it phenomenological} fit is: \begin{eqnarray} \bar{{\cal D}}_{-}^{\it (int)}(X)\ =\ e^{-0.789X^{1/\sigma}}\ \times \nonumber \\ (0.021+0.002X+0.531X^2-0.266X^3+0.261X^4) \label{aval_fit_3d} \end{eqnarray} with $1/\sigma=4.20$ which is obtained from scaling collapses. The distribution curves obtained using the above fit are plotted (thin lines in figure\ \ref{aval_3fig}) alongside the raw data (thick lines). They agree remarkably well even far above $R_c$. We should recall though, from the mean field discussion (see figure\ \ref{mf_aval_fit}), that the fitted curve to the collapsed data can differ from the ``real'' scaling function even for large sizes and close to the critical disorder (in mean field the error was about $10\%$). We expect a similar behavior in finite dimensions. \begin{figure} \centerline{ \psfig{figure=Figures/Tau_Sigma_Beta_Delta_d3_L320_paper_over_R_Rc_legend_THESIS.ps,width=3truein} } \nobreak \centerline{ \psfig{figure=Figures/Sigma_d3_L320_paper_over_R_Rc_legend_THESIS.ps,width=3truein} } \caption[Integrated avalanche size distribution exponents $\tau+\sigma\beta\delta$ and $\sigma$ in $3$ dimensions] {(a) and (b) $\tau+\sigma\beta\delta$ and $\sigma$ respectively, from collapses of the {\bf integrated avalanche size distribution curves} for a $320^3$ spin system. The data is plotted as in mean field. The two closest points to $|r|_{\it avg}=0$ are for a $800^3$ system, for a collapse using curves with disorder: $2.26$, $2.28$, $2.30$, $2.32$, $2.34$, and $2.36$. The extrapolation to $|r|_{\it avg}=0$ gives: $\tau + \sigma \beta \delta = 2.03$ and $\sigma = 0.24$. \label{3d_aval_expfig}} \end{figure} The scaling function in the inset of figure\ \ref{aval_3fig} has a peculiar shape: it grows by a factor of ten before cutting off. The consequence of this shape is that in the simulations, it takes many decades in the size distribution for the slope to converge to the asymptotic power law. This can be seen from the comparison between a straight line fit through the $R=2.25$ ($1000^3$!) curve in figure\ \ref{aval_3fig} and the asymptotic power law $S^{-2.03}$ obtained from scaling collapses and the extrapolation (thick dashed straight line in the same figure). A similar ``bump'' exists in other dimensions and mean field as well. Figure\ \ref{bump_345fig} shows the scaling functions in different dimensions and in mean field. In this graph, the scaling functions are normalized to one and the peaks are aligned (the scaling forms allow this). The curves plotted in figure\ \ref{bump_345fig} are not raw data but fits to the scaling collapse in each dimensions, as was done in the inset of figure\ \ref{aval_3fig}. The mean field and $3$ dimensions curves are given by equations\ (\ref{int_aval3a}) and (\ref{aval_fit_3d}) respectively. For $5$, $4$, and $2$ dimensions, we have respectively: \begin{eqnarray} \bar{{\cal D}}_{-}^{\it (int)}(X)\ =\ e^{-0.518 X^{1/\sigma}}\ \times \nonumber \\ (0.112 + 0.459 X - 0.260 X^2 + 0.201 X^3 - 0.050 X^4) \label{aval_fit_5d} \end{eqnarray} \begin{eqnarray} \bar{{\cal D}}_{-}^{\it (int)}(X)\ =\ e^{-0.954 X^{1/\sigma}}\ \times \nonumber \\ (0.058 + 0.396 X + 0.248 X^2 - 0.140 X^3 + 0.026 X^4) \label{aval_fit_4d} \end{eqnarray} \begin{eqnarray} \bar{{\cal D}}_{-}^{\it (int)}(X)\ =\ e^{-1.076 X^{1/\sigma}}\ \times \nonumber \\ (0.492 - 4.472 X + 14.702 X^2 - \nonumber \\ 20.936 X^3 + 11.303 X^4) \label{aval_fit_2d} \end{eqnarray} with $1/\sigma=2.35, 3.20$, and $10.0$. The errors in the fits are in the same range as for the mean field simulation data (see figure\ \ref{mf_aval_fit}). The $2$ dimensional fit plotted in grey will be covered further in the next section. \begin{figure} \centerline{ \psfig{figure=Figures/curly_D_all_norm_shifted_chopped_paper_THESIS.ps,width=3truein} } \caption[Integrated avalanche size distribution scaling functions in $2$, $3$, $4$, and $5$ dimensions, and mean field] {{\bf Integrated avalanche size distribution scaling functions in $2$, $3$, $4$, and $5$ dimensions, and mean field.} The curves are fits (see text) to the scaling collapses done with exponents from Table\ \protect\ref{measured_exp_table} and\ \protect\ref{conj_meas_2d_table}. The peaks are aligned to fall on (1,1). Due to the ``bump'' in the scaling function the power law exponent can not be extracted from a linear fit to the raw data. \label{bump_345fig}} \end{figure} From figure\ \ref{bump_345fig} we can conclude that in each dimension (and in mean field!), a straight line fit to the integrated avalanche size distribution data is going to give the {\it wrong} critical exponent, and that only by doing scaling collapses and an extrapolation the asymptotic value can be found. This is shown for $3$ dimensions in figure\ \ref{aval_3fig}, and was found to be true in other dimensions as well. We will next see that this is different for the {\it binned} avalanche size distribution. The value for the slope obtained from a linear fit to the data agrees very well with the value obtained from the scaling collapses. \begin{figure} \centerline{ \psfig{figure=Figures/Bin_H_norm_L80_Hc1.265_paper_legend_THESIS.ps,width=3truein} } \nobreak \centerline{ \psfig{figure=Figures/Bin_H_norm_L80_Hc1.265_above_and_below_collapse_paper_THESIS.ps,width=3truein} } \caption[Binned in $H$ avalanche size distribution curves in $4$ dimensions] {(a) {\bf Binned in $H$ avalanche size distribution in $4$ dimensions} for a system of $80^4$ spins at $R=4.09$ ($R_c=4.10$). The critical field is $H_c=1.265$. The curves are averages over close to $60$ random field configuration. Only a few curves are shown. (b) Scaling collapse of the binned avalanche size distribution for $H<H_c$ (upper collapse) and $H>H_c$ (lower collapse). The critical exponents are $\tau=1.53$ and $\sigma\beta\delta=0.54$, and the critical field is $H_c=1.265$. The bins are at fields: $1.162$, $1.185$, $1.204$, $1.220$, $1.234$, $1.245$, $1.254$, $1.276$, $1.285$, $1.296$, $1.310$, $1.326$, $1.345$, and $1.368$. Notice that the two scaling functions do not have a ``bump'' (see text). \label{bin_aval_4dfig}} \end{figure} \paragraph{Binned in $H$ Avalanche Size Distribution} The avalanche size distribution can also be measured at a field $H$ or in a small range of fields centered around $H$. We have measured this ${\it binned}$ in $H$ avalanche size distribution for systems at the critical disorder $R_c$ ($r=0$). To obtain the scaling form, we start from the distribution of avalanches at field $H$ and disorder $R$ (eqn. \ref{int_aval0}): \begin{equation} D(S,R,H) \sim S^{-\tau}\ {\cal D}_{\pm}(S^\sigma |r|, |h|/|r|^{\beta\delta}) \label{aval_distr1} \end{equation} where as before ${\cal D}_{\pm}$ is the scaling function and $\pm$ indicates the sign of $r$. (For most of our data, we can ignore the corrections due to the ``rotation'' of axis as explained in appendix A.) The scaling function can be rewritten as ${\hat {\cal D}}_{\pm}\Bigl(S^\sigma |r|, (S^\sigma |r|)^{\beta\delta} |h|/|r|^{\beta\delta}\Bigr)$, where ${\hat D}_{\pm}$ is a new scaling function. Letting $R \rightarrow R_c$, the scaling for the avalanche size distribution at the field $H$, measured at the critical disorder $R_c$ is: \begin{equation} D(S,H) \sim S^{-\tau}\ {\hat {\cal D}}_{\pm}(|h| S^{\sigma\beta\delta}) \label{aval_distr2} \end{equation} \begin{figure} \centerline{ \psfig{figure=Figures/Tau_norm_d4_L80_paper_legend_new_THESIS.ps,width=3truein} } \caption[Exponent $\tau$ from the binned in $H$ avalanche size distribution curves in $4$ dimensions] {{\bf Values for the exponent $\tau$ extracted from the binned in $H$ avalanche size distribution curves in $4$ dimensions}, for a $80^4$ spin system at $R=4.09$ ($R_c=4.10$). The critical field is $H_c=1.265$. The exponent $\tau$ is found from this linear extrapolation to $\Delta H_{avg} = 0$. The exponent $\sigma\beta\delta$ is calculated from the value of $\tau+\sigma\beta\delta$, extracted from the integrated avalanche size distribution, and the value of $\tau$ from this plot. \label{bin_aval_4dfigc}} \end{figure} Figure\ \ref{bin_aval_4dfig}a shows the binned in $H$ avalanche size distribution curves in $4$ dimensions, for values of $H$ below the critical field $H_c$. (The curves and analysis are similar in $3$ and $5$ dimensions; results in $4$ dimensions are used here for variety.) The simulation was done at the best estimate of the critical disorder $R_c$ ($4.1$ in $4$ dimensions). The binning in $H$ is logarithmic and started from an approximate critical field $H_c$ obtained from the magnetization curves; better estimates of $H_c$ are obtained from the binned distribution data curves and their collapses. Our best estimate for the critical field $H_c$ in $4$ dimensions is $1.265 \pm 0.007$. The scaling form for the logarithmically binned data is the same as in equation\ (\ref{aval_distr2}), if the log-binned data is normalized by the size of the bin. Figure\ \ref{bin_aval_4dfig}b shows the scaling collapse for our data, both below {\it and} above the critical field $H_c$. The ``top'' collapse gives the shape of the ${\hat {\cal D}_{-}}$ ($H<H_c$) function, while the ``bottom'' collapse gives the ${\hat {\cal D}_{+}}$ ($H>H_c$) function. Above the critical field $H_c$, there are spanning avalanches in the system\ \cite{note4}. These are not included in the binned avalanche size distribution collapse shown in figure\ \ref{bin_aval_4dfig}b. The exponent $\tau$ which gives the power law behavior of the binned avalanche size distribution is obtained from an extrapolation similar to previous ones (figure\ \ref{bin_aval_4dfigc}), but with the field $H$ ($\Delta H_{avg}$ in figure\ \ref{bin_aval_4dfigc} is the algebraic average of $H-H_c$ for three curves collapsed together) as the variable instead of the disorder $R$. The exponent $\sigma\beta\delta$ is found to be very sensitive to $H_c$, while $\tau$ is not. We have therefore used the values of $\tau + \sigma\beta\delta$ and $\sigma$ from the integrated avalanche size distribution collapses, and $\tau$ from the binned avalanche size distribution collapses to further constrain $H_c$ (by constraining $\sigma\beta\delta$), and to calculate $\beta\delta$. The latter is then used to obtain collapses of the magnetization curves. We should mention here that $H_c$ in all the dimensions is difficult to find and that it is influenced by finite sizes. The values listed in Table\ \ref{RH_table} are the best estimates obtained from the largest system sizes we have. Nevertheless, systematic errors for $H_c$ could be larger than the errors given in Table\ \ref{RH_table}. This implies possible systematic errors for $\sigma\beta\delta$ which depends on $H_c$, and for $\beta\delta$ which is calculated from $\sigma\beta\delta$. These could also be larger than the errors listed in Table\ \ref{calculated_exp_table}. \begin{figure} \centerline{ \psfig{figure=Figures/Bin_H_norm_L80_Hc1.265_paper_linear_fit_THESIS.ps,width=3truein} } \caption[Linear fit to a binned in $H$ avalanche size distribution curve in $4$ dimensions] {{\bf Binned avalanche size distribution curve (dashed line) in $4$ dimensions}, for a system of $80^4$ spins at $R_c=4.09$. The magnetic field is $H=1.265$. The straight solid line is a linear fit to the data for $S < 13,000$ spins. The slope from the fit is $1.55$ (this varies by not more than $3\%$ as the range over which the fit is done is changed), while the exponent $\tau$ obtained from the collapses and the extrapolation in figure\ \protect\ref{bin_aval_4dfigc} is $1.53 \pm 0.08$. \label{bin_aval_fit_4dfig}} \end{figure} From figure\ \ref{bin_aval_4dfig}b, we see that the two binned avalanche size distribution scaling function do not have a ``bump'' as does the scaling function for the integrated avalanche size distribution (inset in figure \ref{aval_3fig}). Therefore, we expect that the exponent $\tau$ which gives the slope of the distribution in figure\ \ref{bin_aval_4dfig}a can also be obtained by a linear fit through the data curve closest to the critical field. Figure\ \ref{bin_aval_fit_4dfig} shows the curve for the $H=1.265$ bin (dashed curve) as well as the linear fit. The slope from the linear fit is $1.55$ while the value of $\tau$ obtained from the collapses and the extrapolation in figure\ \ref{bin_aval_4dfigc} is $1.53 \pm 0.08$. Fitting the binned distribution curves with a straight line to extract the exponent $\tau$ is also possible in other dimensions and mean field as well. \section{Comparison with the analytical results} We have compared the simulation results with the renormalization group analysis of the same system\ \cite{Dahmen1,Dahmen2}. According to the renormalization group the upper critical dimension (UCD), at and above which the critical exponents are equal to the mean field values, is six. Close to the UCD, it is possible to do a $6-\epsilon$ expansion ($\epsilon$ is small and greater than $0$), and obtain estimates for the critical exponents and the magnetization scaling function, which can then be compared with our numerical results. Furthermore, at dimension eight there is a prediction for another transition. Below eight dimensions, there is a discontinuity in the slope of the magnetization curve as it approaches the ``jump'' in the magnetization ($R < R_c$), while above eight dimensions the approach is smooth. \begin{figure} \centerline{ \psfig{figure=Figures/Exponents_paper_new_THESIS.ps,width=3truein} } \caption[Comparison between the critical exponents from the simulation and the $\epsilon$ expansion] {Numerical values (filled symbols) of the exponents $\tau + \sigma\beta\delta$, $\tau$, $1/\nu$, $\sigma\nu z$, and $\sigma\nu$ (circles, diamond, triangles up, squares, and triangle left) in $2$, $3$, $4$, and $5$ dimensions. The empty symbols are values for these exponents in mean field (dimension 6). Note that the value of $\tau$ in $2$d is the conjectured value: we have not extracted $\tau$ from scaling collapses (see text). We have simulated sizes up to $30000^2$, $1000^3$, $80^4$, and $50^5$, where for $320^3$ for example, more than $700$ different random field configurations were measured. The long-dashed lines are the $\epsilon$ expansions to first order for the exponents $\tau + \sigma\beta\delta$, $\tau$, $\sigma\nu z$, and $\sigma\nu$. The short-dashed lines are Borel sums\protect\cite{LeGuillou-Kleinert} for $1/\nu$. The lowest is the variable-pole Borel sum from LeGuillou {\it et al.}\protect\cite{LeGuillou-Kleinert}, the middle uses the method of Vladimirov {\it et al.} to fifth order, and the upper uses the method of LeGuillou {\it et al.}, but without the pole and with the correct fifth order term. The error bars denote systematic errors in finding the exponents from extrapolation of the values obtained from collapses of curves at different disorders $R$. Statistical errors are smaller. \label{exp_compare}} \end{figure} Figure\ \ref{exp_compare} shows the numerical and analytical results for five of the critical exponents obtained in dimensions two to six (in six dimensions, the values are the mean field ones). The other exponents can be obtained from scaling relations\cite{Dahmen1,Dahmen3}. The exponent values in figure\ \ref{exp_compare} are obtained by extrapolating the results of scaling collapses to either $R \rightarrow R_c$ or $1/L \rightarrow 0$ (see section on simulation results). In two dimensions, which is possibly the lower critical dimension, the plotted values are averages from the three different scaling forms used to collapse the data and extract the exponents. The error bars shown span all three {\it ans\"atze}, and are compatible with our conjectures from the previous section. The long-dashed lines are the $\epsilon$ expansions to first order for $\tau + \sigma\beta\delta$, $\tau$, $\sigma\nu z$, and $\sigma\nu$. The three short-dashed lines\cite{Dahmen1} are Borel sums\protect\cite{LeGuillou-Kleinert} for $1/\nu$. The lowest is the variable-pole Borel sum from LeGuillou {\it et al.}\protect\cite{LeGuillou-Kleinert}, the middle uses the method of Vladimirov {\it et al.} to fifth order, and the upper uses the method of LeGuillou {\it et al.}, but without the pole and with the correct fifth order term\ \cite{Dahmen1}. Notice that the numerical values converge nicely to the mean field predictions, as the dimension approaches six, and that the agreement between the numerical values and the $\epsilon$ expansion is quite impressive. \begin{figure} \centerline{ \psfig{figure=Figures/dMdH_d5_L30_dmdh_epsilon_scaled_paper_THESIS.ps,width=3truein} } \caption[Comparison between simulated $dM/dH$ curves in $5$ dimensions, and the $dM/dH$ curve obtained from the $\epsilon$ expansion] {Comparison between six simulation curves (thin lines) and the $dM/dH$ curve (thick dashed line) obtained from a parametric form\ \protect\cite{Zinn-Justin} to third order in $\epsilon$. The six curves are for a system of $30^5$ spins at disorders: $7.0, 7.3,$ and $7.5$ ($R_c=5.96$ in $5$ dimensions), and for a system of $50^5$ spins at disorders: $6.3, 6.4$, and $6.5$ (for larger fields, these are closer to the dashed line in the figure). All the curves have been stretched/shrunk in the horizontal and vertical direction to lie on each other, and shifted horizontally. \label{5d_dmdh_fig}} \end{figure} The $\epsilon$ expansion can be an even more powerful tool if it can predict the scaling functions. This has been done for the magnetization scaling function of the pure Ising model in $4-\epsilon$ dimensions\ \cite{DombWallace,Zinn-Justin}. Since the $\epsilon$ expansion for our model is the same as the one for the {\it equilibrium} RFIM\ \cite{Dahmen1}, and the latter has been mapped to {\it all} orders in $\epsilon$ to the corresponding expansion of the regular Ising model in two lower dimensions\ \cite{Dahmen1,Aharony,Parisi}, we can use the results obtained in\ \cite{DombWallace,Zinn-Justin}. This is done in figure\ \ref{5d_dmdh_fig}, which shows the comparison between the $dM/dH$ curves obtained in $5$ dimensions at $R=6.3, 6.4, 6.5, 7.0, 7.3, 7.5$ ($R_c=5.96$) (the curves have been stretched/shrunk to lie on top of each other, and shifted horizontally so that the peaks align), and the parametric form (thick dashed line) for the scaling function of $dM/dH$, to third order in $\epsilon$, where $\epsilon =1$ in $5$ dimensions (see\ \cite{Zinn-Justin}). As we see, the agreement is very good in the scaling region (close to the peaks).This brings up the possibility of using the $\epsilon$ expansion for the scaling function to extract the critical exponents from simulation or experimental data. So far though, only the scaling function for the magnetization has been obtained. \begin{figure} \centerline{ \psfig{figure=Figures/MofH_579d_paper_THESIS.ps,width=3truein} } \caption[Magnetization curves showing the approach to the ``infinite jump'' in $5$, $7$, and $9$ dimensions] {{\bf Magnetization curves in $5$, $7$, and $9$ dimensions.} The disorders for these curves are $R=3.3$, $4.7$, and $6.0$ for $30^5$, $10^7$, and $5^9$ size systems respectively. The dashed lines represent the ``jump'' in the magnetization. Notice that in $9$ dimensions the approach to the ``jump'' seems to be continuous. \label{transition_8d}} \end{figure} As another check between the simulation and the renormalization group, we have looked for the predicted transition in eight dimensions. Figure\ \ref{transition_8d} shows the magnetization curves in $5$, $7$, and $9$ dimensions (system sizes: $30^5$, $10^7$, and $5^9$) for values of the disorder equal to ${2 \over 3}d$, where $d$ is the dimension. These values of disorder are below the critical disorder in dimensions below six, and are expected to be below for dimensions $7$ and $9$ as well. For $5$d and $7$d, the approach to the ``jump'' in the magnetization is discontinuous. Above the eight dimension, the approach is continuous (see close ups in figure\ \ref{closeup_8d}). This is as expected from the renormalization group\ \cite{Dahmen1}. We have also looked at $dM/dH$, which appears clearly to diverge in $d=9$ and not in $d=7$ (figure\ \ref{slope_8d}). \section{Conclusion} We have used the zero temperature random field Ising model, with a Gaussian distribution of random fields, to model a random system that exhibits hysteresis. We found that the model has a transition in the shape of the hysteresis loop, and that the transition is critical. The tunable parameters are the amount of disorder $R$ and the external magnetic field $H$. The transition is marked by the appearance of an infinite avalanche in the thermodynamic system. Near the critical point, ($R_C$, $H_C$), the scaling region is quite large: the system can exhibit power law behavior for several decades, and still not be near the critical transition. This is important to keep in mind whenever experimental data are analyzed. If a tunable parameter can be found, a system that appears to be SOC, might in reality have a disorder induced critical point. \begin{figure} \centerline{ \psfig{figure=Figures/MofH_79d_closeup_paper_THESIS.ps,width=3truein} } \caption[Closeup of magnetization curves with the ``infinite jump'' in $7$ and $9$ dimensions] {(a) and (b) Closeup of the {\bf magnetization curves in $7$ and $9$ dimensions} respectively from figure\ \protect\ref{transition_8d}. In $8$ dimensions, there is a prediction from the renormalization group\ \protect\cite{Dahmen1} that there is a transition in the way the jump is approached (see text). \label{closeup_8d}} \end{figure} We have extacted critical exponents for the magnetization, the avalanche size distribution (integrated over the field and binned in the field), the moments of the avalanche size distribution, the avalanche correlation, the number of spanning avalanches, and the distribution of times for different avalanche sizes. These values are listed in Table\ \ref{measured_exp_table} and Table\ \ref{conj_meas_2d_table}, and were obtained as an average of the extrapolation results (to $R \rightarrow R_c$ or $L \rightarrow \infty$) from several measurements. For example, the correlation length exponent $\nu$ is the average value from three different collapses: the correlation function, the spanning avalanches, and the second moments of the avalanche size distribution, while the critical disorder $R_c$ is estimated from both the spanning avalanches collapses and the collapses of the moments of the avalanche size distribution. As shown earlier, the numerical results compare well with the $\epsilon$ expansion\ \cite{Dahmen1,Dahmen2}. Furthermore, the renormalization group work predicts another transition in eight dimensions, which we find in the simulation as well. Comparisons to experimental Barkhausen noise measurements\ \cite{Perkovic} are very encouraging, and a more comprehensive review of possible experiments that exhibit disorder--driven critical phenomena similar to our model is under way\ \cite{Dahmen3}. \begin{figure} \centerline{ \psfig{figure=Figures/dMdH_579d_paper_THESIS.ps,width=3truein} } \caption[$dM/dH$ for magnetization curves with the ``infinite jump'', in $7$ and $9$ dimensions] {Derivative of the magnetization with respect to the field $H$, for the curves in figure\ \protect\ref{transition_8d}. The approach to the ``infinite jump'' seems to be continuous in $9$ dimensions. Note that the vertical axis is logarithmic. \label{slope_8d}} \end{figure} Finally, we should mention that there are other models for avalanches in disordered magnets. There is a large body of work on depinning transitions and the motion of the single interface\cite{depinning,SameDepinning}. In these models, avalanches occur only at the growing interface. Our model though, deals with many interacting interfaces: avalanches can grow anywhere in the system. Similar models exist with random bonds\cite{RandomBonds,Vives} and random anisotropies. In the random bonds model, the interaction $J_{ij}$ between neighboring spins $i$ and $j$ is random. The zero temperature random bond Ising model\ \cite{RandomBonds,Vives} also exhibits a critical transition in the shape of the hysteresis loop, where the mean bond strength is analogous to our disorder $R$. It has been argued numerically\ \cite{Vives} and analytically\ \cite{Dahmen1}, that as long as there are no long-range forces\cite{Urbach} and correlated disorder, the random bond and the random field Ising model are in the same universality class. A comparison between our simulation and the results in reference\ \cite{Vives} show that the $3$ dimensional results agree quite nicely. However, in $2$ dimensions, there are large differences, which we believe occur because of the small system sizes used by the authors for their simulation (only up to $L=100$). We have seen that our results (see section on the $2$ dimensional simulation) are very size dependent. Looking back for example at figure\ \ref{span_2d_fig}, we find that for a system of $L=100$ spins, a ``good'' estimate for the critical disorder $R_c$ would indeed be $0.75$ as was found in\ \cite{Vives}. However, we find after increasing the system size that the critical disorder $R_c$ is $0.54$ or lower. \subsubsection{Avalanche Correlation} The avalanche correlation function $G(x,R,H)$ measures the probability that a flipping spin will trigger, through an avalanche of spins, another spin a distance $x$ away. From the renormalization group description\cite{Dahmen1,Dahmen2}, close to the critical point and for large distances $x$, the correlation function is given by (corrections are subdominant as explained in appendix A): \begin{equation} G(x,R,H) \sim {1 \over {x^{d-2+\eta}}}\ {\cal G}_{\pm}(x/{\xi(r,h)}) \label{correl_equ1} \end{equation} where $r$ and $h$ are respectively the reduced disorder and field, ${\cal G_{\pm}}$ ($\pm$ indicates the sign of $r$) is the scaling function, $d$ is the dimension, $\xi$ is the correlation length, and $\eta$ is called the ``anomalous dimension''. The correlation length $\xi (r,h)$ is a macroscopic length scale in the system which is roughly on the order of the mean linear extent of the avalanches for a system away from the critical point. \begin{figure} \centerline{ \psfig{figure=Figures/Norm_correl_d3_L320_paper_THESIS.ps,width=3truein} } \caption[Avalanche correlation curves in $3$ dimensions] {{\bf Avalanche correlation function integrated over the field $H$ in $3$ dimensions}, for $L=320$. The curves are averages of up to $19$ random field configurations. The critical disorder $R_c$ is $2.16$. \label{correl_3d_fig}} \end{figure} {\noindent At the critical field $H_c$ (h=0) and near $R_c$, the correlation length scales like $\xi \sim |r|^{-\nu}$, while for small field $h$ it is given by $\xi \sim |r|^{-\nu}\ {\cal Y}_{\pm}(h/|r|^{\beta\delta})$ where ${\cal Y}_{\pm}$ is a universal scaling function. The avalanche correlation function should not be confused with the cluster or ``spin-spin'' correlation which measures the probability that two spins a distance $x$ away have the same value. (The algebraic decay for this other, spin-spin correlation function at the critical point ($r=0$ and $h=0$), is $1/{x^{d-4+{\tilde \eta}}}$\cite{Dahmen1}.)} \begin{figure} \centerline{ \psfig{figure=Figures/Norm_correl_d3_L320_collapse_paper_THESIS.ps,width=3truein} } \nobreak \nobreak \centerline{ \psfig{figure=Figures/Nu_d3_L320_correl_paper_THESIS.ps,width=3truein} } \caption[Scaling collapse of the avalanche correlation curves in $3$ dimensions, and exponent $\nu$] {(a) Scaling collapse of the {\bf avalanche correlation function integrated over the field $H$, in $3$ dimensions} for $L=320$. The values of the disorders range from $R=2.35$ to $R=3.0$, with $R_c=2.16$. The exponents are: $\nu=1.39 \pm 0.20$ and $d + \beta/\nu = 3.07 \pm 0.30$. (b) Exponent $\nu$ extracted from collapses of avalanche correlation curves (see (a)). The extrapolated value at $|r|_{avg}=0$ is $1.37 \pm 0.18$. \label{correl_collapse_3d_fig}} \end{figure} We have measured the avalanche correlation function integrated over the field $H$, for $R>R_c$. For every avalanche that occurs between $H=-\infty$ and $H =+\infty$, we keep a count on the number of times a distance $x$ occurs in the avalanche. To decrease the computational time not every pair of spins is selected; instead we do a statistical average for $S$ pairs where $S$ is the size of the avalanche. Our simulation seems to indicate that the difference between this statistical average and the exact measurement is less than the fluctuations obtained from measurements of the correlation function for different realizations of the random field distribution. The data is saved in ``distance'' bins separated by $0.5$ and starting at a distance of $1.0$ (the self correlation is not included), and is normalized by the number of neighbors at each distance. The spanning avalanches are not included in our correlation measurement. Figure\ \ref{correl_3d_fig} shows several avalanche correlation curves in $3$ dimensions for $L=320$. The scaling form for the avalanche correlation function integrated over the field $H$, close to the critical point and for large distances $x$, is obtained by integrating equation (\ref{correl_equ1}): \begin{equation} G_{\it int}(x,R) \sim \int {1 \over {x^{d-2+\eta}}}\ {\cal G}_{\pm}\Bigl(x/{\xi(r,h)}\Bigr)\ dh \label{correl_equ2} \end{equation} Near the critical point $\xi(r,h) \sim |r|^{-\nu} {\cal Y}_{\pm} (h/|r|^{\beta\delta})$. Defining $u=h/|r|^{\beta\delta}$, equation (\ref{correl_equ2}) becomes: \begin{equation} G_{\it int}(x,R) \sim |r|^{\beta\delta}{x^{-(d-2+\eta)}} \int {\cal G}_{\pm}\Bigl(x/|r|^{-\nu} {\cal Y}_{\pm} (u)\Bigr)\ du \label{correl_equ3} \end{equation} The integral ($\cal I$) in equation (\ref{correl_equ3}) is a function of $x|r|^{\nu}$ and can be written as: \begin{equation} {\cal I} = (x|r|^{\nu})^{-\beta\delta/\nu}\ {\widetilde {\cal G}}_{\pm}(x|r|^{\nu}) \label{correl_equ4} \end{equation} to obtain the scaling form: \begin{equation} G_{\it int}(x,R) \sim {1 \over x^{d+\beta/\nu}}\ {\widetilde {\cal G}}_{\pm}(x|r|^{\nu}) \label{correl_equ5} \end{equation} where we have used the scaling relation $(2-\eta)\nu=\beta\delta-\beta$ (see\ \cite{Dahmen1,Dahmen3} for the derivation). \begin{figure} \centerline{ \psfig{figure=Figures/Norm_Correl_and_Aniso_d3_L320_paper_THESIS.ps,width=3truein} } \caption[Avalanche correlation function anisotropies in $3$ dimensions] {{\bf Anisotropies in the avalanche correlation function}. The curves are for a system of $320^3$ spins at $R=2.35$. Four curves are shown on the graph: one is the avalanche correlation function integrated over the field $H$ (as in figure\ \protect\ref{correl_3d_fig}), while the other three are measurements of the correlation along the three axis, the six face diagonals, and the four body diagonals. Avalanches involving more than four spins show no noticeable anisotropy: the critical point appears to have spherical symmetry. The same result is found in $2$ dimensions. \label{correl_aniso_3d_fig}} \end{figure} Figure\ \ref{correl_collapse_3d_fig}a shows the integrated avalanche correlation curves collapse in $3$ dimensions for $L=320$ and $R>R_c$. The exponent $\nu$ is obtained from such collapses by extrapolating to $R = R_c$ (figure\ \ref{correl_collapse_3d_fig}b) as was done for other collapses. The exponent $\beta/\nu$ can be obtained from these collapses too, but it is much better estimated from the magnetization discontinuity covered below. The value of $\beta/\nu$, listed in Table\ \ref{measured_exp_table} alongside all the other exponents, is derived from the magnetization discontinuity collapses only. We have also looked for possible anisotropies in the integrated avalanche correlation function in $2$ and $3$ dimensions. The anisotropic integrated avalanche correlation functions are measured along ``generalized diagonals'': one along the three axis, the second along the six face diagonals, and the third along the four body diagonals. We compare the integrated avalanche correlation function and the anisotropic integrated avalanche correlation functions to each other, and find no anisotropies in the correlation, as can be seen from figure\ \ref{correl_aniso_3d_fig}. \section{Introduction} The increased interest in real materials in condensed matter physics has brought disordered systems into the spotlight. Dirt changes the free energy landscape of a system, and can introduce metastable states with large energy barriers\ \cite{barrier}. This can lead to extremely slow relaxation towards the equilibrium state. On long length scales and practical time scales, a system driven by an external field will move from one metastable local free-energy minimum to the next. The equilibrium, global free energy minimum and the thermal fluctuations that drive the system toward it, are in this case irrelevant. The state of the system will instead depend on its history. The motion from one local minima to the next is a collective process involving many local (magnetic) domains in a local region - {\it an avalanche}. In magnetic materials, as the external magnetic field $H$ is changed continuously, these avalanches lead to the magnetic noise: the Barkhausen effect\cite{Jiles,McClure}. This effect can be picked up as voltage pulses in a coil surrounding the magnet. The distribution of pulse (avalanche) sizes is found\ \cite{McClure,Barkhausen,Urbach,SOC_example} to follow a power law with a cutoff after a few decades, and was interpreted by some\ \cite{SOC_example} to be an example of self-organized criticality\ \cite{SOC}. (In SOC, a system organizes itself into a critical state without the need to tune an external parameter.) Other systems can exhibit avalanches as well. Several examples where disorder may play a part are: superconducting vortex line avalanches\ \cite{Field}, resistance avalanches in superconducting films\ \cite{Wu}, and capillary condensation of helium in Nuclepore\ \cite{Nuclepore}. The history dependence of the state of the system leads to hysteresis. Experiments with magnetic tapes\ \cite{Berger} have shown that the shape of the hysteresis curve changes with the annealing temperature. The hysteresis curve goes from smooth to discontinuous as the annealing temperature is increased. This transition can be explained in terms of a {\it plain old critical point} with two tunable parameters: the annealing temperature and the external field. At the critical temperature and field, the correlation length diverges, and the distribution of pulse (avalanche) sizes follows a power law. We have argued earlier\ \cite{Perkovic} that the Barkhausen noise experiments can be quantitatively explained by a model\ \cite{Sethna} with two tunable parameters (external field and disorder), which exhibits {\it universal}, non-equilibrium collective behavior. The model is athermal and incorporates collective behavior through nearest neighbor interactions. The role of {\it dirt} or disorder, as we call it, is played by random fields. This paper presents the results and conclusions of a large scale simulation of that model: the non-equilibrium zero-temperature Random Field Ising Model (RFIM), with a deterministic dynamics. The results compare very well to our $\epsilon$ expansion\cite{Dahmen1,Dahmen2}, and to experiments in Barkhausen noise\cite{Perkovic}. A more detailed comparison to experimental systems is in process\cite{Dahmen3}. The paper is divided as follows. Section II quickly reviews the model. Section III explains the simulation method that we use. Section IV explains the data analysis and shows results for the simulation in $2$, $3$, $4$, and $5$ dimensions, as well as in mean field. Section V gives a comparison between the simulation and the $\epsilon$ expansion exponents, and a comparison between the shape of the magnetization curves in $5$, $7$, and $9$ dimensions, and the predicted shape from the $\epsilon$ expansion. Section VI summarizes the results. This is followed by three appendices that cover derivations that were omitted in the text for continuity. \subsection{Mean Field Simulation} The mean field simulation shows how well the results for the critical exponents, obtained close to $R_c$ and for finite size systems (finite number of spins), extrapolate to the calculated values for a system in the thermodynamic limit, at the critical disorder. Thus, we will omit in this section some details that are only relevant for understanding the non-mean field simulation results. We start with the curves for the magnetization as a function of the field for different values of the disorder, which we find are not useful for extracting critical exponents. We then go on to measurements of spin avalanche sizes and their moments. Avalanches that span the system from one ``side'' to another will also be mentioned although since in mean field there are no ``sides'', we will define what we mean by a mean field spanning avalanche. Since distances are irrelevant in mean field, we do not have any correlation measurements, but we can still apply what we learn from other collapses in mean field to the correlation measurement data in $2$, $3$, $4$, and $5$ dimensions. Figure\ \ref{mf_mofhfiga} shows the magnetization curves, and figure \ref{mf_mofhfig}a shows a scaling collapse for a $10^6$ mean field spin system and $r<0$ ($R>R_c$). As mentioned earlier, near the critical point ($R_c = {\sqrt {2/\pi}}$ for $J=1$, in mean field), the magnetization scales like\cite{Sethna,Dahmen1} \begin{equation} M(H,R) - M_c(H_c,R_c) \sim |r|^\beta\ {\cal M}_{\pm}(h/|r|^{\beta\delta}) \label{mf_mofh_equ1} \end{equation} where $\pm$ refers to the sign of the reduced disorder $r=(R_c-R)/R$, and $h=(H-H_c)$. The mean field critical exponents are $\beta = 1/2$ and $\beta\delta = 3/2$. Notice in figure\ \ref{mf_mofhfig}a that the scaling region around $M_c=0$ and $H_c=0$ is very small; figure\ \ref{mf_mofhfig}b shows that a substantially different set of critical exponents leads to a similar looking collapse. In general, the critical field $H_c$ and the critical magnetization $M_c$ are not zero as in mean field, and $M_c$ is not well determined numerically. In dimensions that we simulate ($2$ through $5$), the critical region is not only small but it is also poorly defined, which does not sufficiently constrain the values of the exponents. This makes the magnetization function $M(H,R)$ a poor choice for extracting critical exponents. \begin{figure} \centerline{ \psfig{figure=Figures/MofH_S1000000_all_paper_THESIS.ps,width=3truein} } {\caption[Mean field magnetization curves] {{\bf Mean field magnetization} curves for $10^6$ spins. The critical disorder is $R_c= 0.79788456$. The curves are averages of $6$ to $10$ different initial realizations of the random field distribution. \label{mf_mofhfiga}}} \end{figure} The critical magnetization $M_c$ can be removed from the scaling form if we look at the first derivative of the magnetization with respect to the field instead. $dM/dH$ scales like: \begin{equation} {dM \over dH} (H,R) \sim |r|^{\beta-\beta\delta}\ \dot{\cal M}_{\pm}(h/|r|^{\beta\delta}) \label{dmdh_mf_equ1} \end{equation} where $\dot{\cal M}_{\pm}$ denotes the derivative of the scaling function ${\cal M}_{\pm}$ with respect to its argument $h/|r|^{\beta\delta}$. The $dM/dH$ mean field curves and collapses are shown in figure~\ref{mf_dmdh_figa} and figures~\ref{mf_dmdh_fig}(a--b). Notice that the incorrect exponents $\beta=0.4$ and $\beta\delta=1.65$ give a better collapse (fig.\ \ref{mf_dmdh_fig}b). Figure\ \ref{mf_dmdh_figd} shows a close up of figure\ \ref{mf_dmdh_fig}a, alongside with three (thin) curves for disorders: $0.80,0.81,$ and $0.82$. These are not measured in the simulation (the finite number of mean field spins we use give rise to finite size effects near $R_c$ as we will see soon); instead they are numerically calculated from the mean field implicit equation for the magnetization\ \cite{Sethna,Dahmen1}: \begin{equation} M(H) = 1 - 2 \int_{-\infty}^{-J^{*}M(H)-H} \rho(f)\ df \label{mofh_implicit} \end{equation} where $J^{*}$ denotes the coupling of one spin to {\it all} the other spins in the system, and $\rho(f)$ is the random field distribution function. \begin{figure} \centerline{ \psfig{figure=Figures/MofH_S1000000_collapse_R_all_paper_THESIS.ps,width=3truein} } \nobreak \centerline{ \psfig{figure=Figures/MofH_S1000000_collapse_R_all_paper_1.4_THESIS.ps,width=3truein} } \caption[Scaling collapse of mean field magnetization curves] {(a) {\bf Scaling collapse for the mean field magnetization curves} at disorders $R=0.912$, $0.974$, $1.069$, $1.165$, $1.197$, and $1.460$. (These values of disorder were chosen relative to $R_c=0.79788456$, to match some of the values we measured in $3$ dimensions (see figure\ \protect\ref{3d_MofH_fig}). The value of the critical disorder $R_c$ in $3$ dimensions has since been modified, and there is no correspondence anymore.) The exponents are $\beta=1/2$ and $\beta\delta=3/2$. $m$ is defined as $M-M_c$, and in mean field both $M_c$ and $H_c$ are zero. The inset shows a closeup of the critical region. (b) Scaling collapse of the same curves as in (a) but with the (wrong) exponents $\beta=0.4$ and $\beta\delta =1.65$. The two collapses are very similar. The inset is a closeup. \label{mf_mofhfig}} \end{figure} The scaling collapse converges to the expected scaling function (dashed thick line) as we get closer to the critical disorder. The expected scaling form is also obtained from an analytic expression derived in mean field\ \cite{Sethna,Dahmen1}. It is given by the smallest real root $g(y)$ of the cubic equation: \begin{equation} g^3 + {12 \over \pi} g - {12 \sqrt{2} \over \pi^{3/2} R_c} y = 0. \label{g_eqn} \end{equation} \begin{figure} \centerline{ \psfig{figure=Figures/dMdH_S1000000_less2_points_paper_THESIS.ps,width=3truein} } \caption[Mean field $dM/dH$ curves]{{\bf Mean field $dM/dH$ curves} for $10^6$ spins and disorders $R=0.912$, $0.974$, and $1.069$ (from largest to smallest peak). The original data is the same as in figure\ \protect\ref{mf_mofhfiga}. The critical disorder is $R_c = 0.79788456$. \label{mf_dmdh_figa}} \end{figure} We again find that the critical exponents and $R_c$, obtained from the $dM/dH$ curves, are ill-determined. In finite dimensions, that is even more true since we have another parameter to fit: $H_c$. For dimensions $3$, $4$, and $5$, we extract $\beta$, $\beta\delta$, $H_c$, and $R_c$ by other means and simply show the resulting collapse of the $M(H)$ and $dM/dH$ curves as a check. As mentioned earlier, the spins flip in avalanches of varying sizes. The distribution of ${\it all}$ the avalanches that occur at a disorders $R$ while the external field $H$ is raised adiabatically from $-\infty$ to $+\infty$ is plotted in figure\ \ref{mf_aval_collapfigaa}. The curves in this plot are normalized by the number of spins in the system, and therefore represent the probability {\it per spin} for an avalanche of size $S$ to occur in the hysteresis loop, at disorder $R$. The curves can be normalized to one if they are divided by the total number of avalanches in the loop, and multiplied by the number of spins in the system. Often in experiments, the $\it binned$ avalanche size distribution, which contains only avalanches that occur in a small range of fields around a particular value of the field $H$, is measured instead. The scaling form for this distribution\cite{note2} is\ \cite{Sethna,Dahmen1}: \begin{equation} D(S,R,H) \sim S^{-\tau}\ {\bar {\cal D}_{\pm}} (S^\sigma |r|, h/|r|^{\beta\delta}) \label{int_aval0} \end{equation} where $S$ is the size of the avalanche and is large, and $r$ and $h$ are small. In mean field, $\sigma=1/2$ and $\tau=3/2$. The scaling form for the integrated avalanche size distribution is obtained by integrating the above form over all fields: \begin{equation} D_{\it int}(S,R) \sim \int S^{-\tau}\ {\bar {\cal D}_{\pm}} (S^\sigma |r|, h/|r|^{\beta\delta})\ dh \label{int_aval1} \end{equation} \begin{figure} \centerline{ \psfig{figure=Figures/dMdH_S1000000_less2_points_smooth5_collapse_paper_THESIS.ps,width=3truein} } \nobreak \centerline{ \psfig{figure=Figures/dMdH_S1000000_less2_points_smooth5_collapse_paper_good_THESIS.ps,width=3truein} } \caption[Scaling collapse of mean field $dM/dH$ curves] {(a) {\bf Scaling collapse of mean field $dM/dH$ curves} from figure \protect\ref{mf_dmdh_figa}. The exponents are $\beta=1/2$ and $\beta\delta=3/2$ (mean field values). The curves are smoothed over $5$ data points (using a running average) to show the collapse better. The curves do not collapse well for large and small $h/r^{\beta\delta}$, unless we get very close to the critical disorder (see figure \protect\ref{mf_dmdh_figd}). (b) Scaling collapse of data in (a) but with exponents $\beta=0.4$ and $\beta\delta=1.65$. The collapse is better, although the exponents are wrong. \label{mf_dmdh_fig}} \end{figure} With the change of variable $u=h/|r|^{\beta\delta}$, equation\ (\ref{int_aval1}) becomes: \begin{equation} D_{\it int}(S,R) \sim S^{-\tau}\ |r|^{\beta\delta} \int {\bar {\cal D}_{\pm}} (S^\sigma |r|,u)\ du \label{int_aval2} \end{equation} The integral in equation\ (\ref{int_aval2}) is a function of $S^\sigma |r|$ only,so we can write it as: \begin{equation} (S^\sigma |r|)^{-\beta\delta}\ {\bar {\cal D}_{\pm}}^{(int)} (S^\sigma |r|) \label{int_aval2a} \end{equation} to obtain the scaling form for the integrated avalanche size distribution: \begin{equation} D_{\it int} (S,R) \sim S^{-(\tau + \sigma\beta\delta)}\ {\bar {\cal D}_{\pm}}^{(int)} (S^\sigma |r|) \label{int_aval3} \end{equation} To obtain equation\ (\ref{int_aval2a}), we have assumed that the integral in (\ref{int_aval2}) converges. This is usually safe to do since the distribution curves near the critical point drop off exponentially for large arguments. The same kind of argument can be made for the integrals of other measurements as well. \begin{figure} \centerline{ \psfig{figure=Figures/dMdH_S1000000_less2_points_smooth5_meanfield_collapse_paper_THESIS.ps,width=3truein} } \caption[Close--up of mean field $dM/dH$ curves collapse] {{\bf Close-up of the mean field $dM/dH$ curves collapse} in figure \protect\ref{mf_dmdh_fig}a. Also plotted are three curves (thin lines) calculated using the mean field analytic solution to $M(H)$ (see text). These are for $R=0.80$, $0.81$, and $0.82$. We see that the scaling collapse, at the mean field exponents, of the $dM/dH$ curves converges to the expected mean field scaling function (thick dashed line), as $R \rightarrow R_c$. \label{mf_dmdh_figd}} \end{figure} \begin{figure} \centerline{ \psfig{figure=Figures/Aval_histo_S1000000_paper_norm_THESIS.ps,width=3truein} } \caption[Mean field integrated avalanche size distribution curves] {{\bf Mean field integrated avalanche size distribution curves} for $10^6$ spins and disorders $R=0.912$, $0.974$, $1.069$, $1.197$, and $1.460$ (from right to left). The straight line is the slope of the power law behavior in mean field: $\tau + \sigma\beta\delta=9/4$. \label{mf_aval_collapfigaa}} \end{figure} Figures\ \ref{mf_aval_collapfiga}a and \ref{mf_aval_collapfiga}b show two collapses with different critical exponents of the curves from figure\ \ref{mf_aval_collapfigaa}, using the scaling form in equation\ (\ref{int_aval3}). Notice that the collapse with the incorrect exponents $\tau + \sigma\beta\delta =2.4$ and $\sigma = 0.44$ is better than the collapse with the mean field exponents $\tau + \sigma\beta\delta = 9/4$ and $\sigma = 1/2$. Although the distribution curves in figures\ \ref{mf_aval_collapfiga}a and \ref{mf_aval_collapfiga}b have disorders that are far from the critical disorder $R_c=0.79788456$, the curves collapse but with the {\it wrong} exponents. \begin{figure} \centerline{ \psfig{figure=Figures/Aval_histo_S1000000_collapse_sigma_R_paper_norm_THESIS.ps,width=3truein} } \nobreak \centerline{ \psfig{figure=Figures/Aval_histo_S1000000_collapse_sigma_R_paper_norm_good_THESIS.ps,width=3truein} } \caption[Scaling collapse of the mean field integrated avalanche size distribution curves] {(a) {\bf Scaling collapse of three integrated avalanche size distribution curves in mean field}, for disorders: $1.069$, $1.197$, and $1.460$. The curves are smoothed over $5$ data points before they are collapsed. The collapse is done using the mean field values of the exponents $\sigma$ and $\tau + \sigma\beta\delta$ ($1/2$ and $9/4$ respectively), and $r = (R_c-R)/R$. (b) Same curves and scaling form as in (a), but with the exponents $\sigma=0.44$ and $\tau + \sigma\beta\delta = 2.4$. The collapse is better for the incorrect exponents! We use this ``best'' collapse to extract exponents for figures\ \protect\ref{mf_aval_expfig}a and \protect\ref{mf_aval_expfig}b, and then extrapolate to $R=R_c$ to obtain the correct mean field exponents. \label{mf_aval_collapfiga}} \end{figure} It is surprising that these curves collapse at all since the scaling form is correct only for $R$ close to $R_c$. Corrections to scaling become important away from the critical point, but it seems that the scaling form has enough ``freedom'' that collapses are possible even far from $R_c$. In the limit of $R \rightarrow R_c$, we expect that the exponents obtained from such collapses will converge to the mean field value, and that the extrapolation will remove the question of scaling corrections. To test this, we have collapsed three curves at a time, and plotted the values of the exponents extracted from such collapses against the average of the reduced disorder $|r|$ for the three curves, which we call $|r|_{\it avg}$ (figures\ \ref{mf_aval_expfig}a and \ref{mf_aval_expfig}b). In these figures, notice two things. First, the linear extrapolation to $|r|_{\it avg}=0$ agrees quite well with the mean field exponent values, and second, the points obtained by doing collapses using $r=(R_c-R)/R$ either converge faster to the mean field exponents or do as well as the points obtained from collapses done with $r=(R_c-R)/R_c$. This is true for all the extrapolations that we have done in mean field. Other models (see for example\ \cite{Robbins}) exhibit this behavior, and experimentalists seem to have known about this for a while\ \cite{Souletie}. \begin{figure} \centerline{ \psfig{figure=Figures/Tau_sigma_beta_delta_S1000000_paper_over_R_Rc_legend_THESIS.ps,width=3truein} } \nobreak \centerline{ \psfig{figure=Figures/Sigma_S1000000_paper_over_R_Rc_legend_THESIS.ps,width=3truein} } \caption[Mean field exponents $\tau + \sigma\beta\delta$ and $\sigma$, from the integrated avalanche size distribution] {(a) $\tau + \sigma \beta \delta$ from collapses of {\bf mean field integrated avalanche size distribution curves} for $10^6$ spins. The two points closest to $|r|_{\it avg}=0$ are for a system of $10^7$ spins. $|r|_{\it avg}$ is the average reduced disorder $|r|$ for the three curves collapsed together (see text). (b) $\sigma$ from collapses of integrated avalanche size distribution curves for the system in (a). Again, the two closest points to $|r|_{\it avg}=0$ are for a system of $10^7$ spins. The mean field values are calculated analytically. \label{mf_aval_expfig}} \end{figure} In dimensions $2$ to $5$, we obtain the exponents $\tau+\sigma\beta\delta$ and $\sigma$ in the limit $R \rightarrow R_c$, using the above linear extrapolation method. For other collapses, if the two extrapolation results differ substantially, we ``bias'' our result towards the $r=(R_c-R)/R$ extrapolated value of the exponent. Notice in figures\ \ref{mf_aval_collapfiga}a and \ref{mf_aval_collapfiga}b that the scaling function ${\bar {\cal D}_{-}^{(int)}}$ has a ``bump'' (the $-$ sign indicates that the collapse is for curves with $R > R_c$). Although we will come back to this point when we talk about the results in $5$ and lower dimensions, it is interesting to know what the shape of the scaling function ${\bar {\cal D}}_{-}^{(int)}$ is. In appendix B, we calculate the mean field scaling function for $r<0$ (equations\ \ref{apA1_eq14} and \ref{apA1_eq15}): \begin{eqnarray} {\bar {\cal D}}_{-}^{(int)}(-r S^{\sigma})\ =\ {e^{-(-r S^{\sigma})^2 \over 2} \over \pi {\sqrt 2}}\ \times \nonumber \\ \int_0^{\infty} e^{\Bigl(-(-r) S^{\sigma}\ u - {u^2 \over 2}\Bigr)}\ \Bigl(-r S^{\sigma} + u\Bigr)\ {du \over {\sqrt u}} \label{int_aval3b} \end{eqnarray} where $\sigma=1/2$. \begin{figure} \centerline{ \psfig{figure=Figures/Aval_histo_mf_fit_paper_THESIS.ps,width=3truein} } \caption[Mean field scaling function for the integrated avalanche size distribution] {Scaling collapse of the {\bf mean field integrated avalanche size distribution curves} (dashed lines), for $S=10^6$ spins and $R=0.912,0.974$, and $S=10^7$ spins and $R=0.854, 0.878, 0.912$. The critical exponents are: $\tau+\sigma\beta\delta =9/4$ and $\sigma =1/2$. The thick black line is the best fit to the data using a function that is the product of a polynomial and an exponential (eqn.\ (\protect\ref{int_aval3a})). The thick grey line is the ``real'' mean field scaling function (see text).\label{mf_aval_fit}} \end{figure} A closed analytic form can not be obtained, but we can find the behavior of this function for small and large arguments $-rS^\sigma$. For small arguments $X=-rS^\sigma$, the scaling function is a polynomial in $X$ (\ref{apA1_eq17}), while for large arguments, the scaling function is given by the product of an exponential decay in $X^2$ and the square root of $X$ (\ref{apA1_eq18}). We can then try to fit our data (the scaling collapse) with a function that will incorporate a polynomial and an exponential decay (as an approximation to the real function). We obtain: \begin{eqnarray} e^{-{X^2 \over 2}}\ & (0.204 + 0.482 X - 0.391 X^2 + \nonumber \\ & 0.204 X^3 - 0.048 X^4) \label{int_aval3a} \end{eqnarray} This form has the expected exponential behavior at large $X$, but the wrong pre-factor. On the other hand, for small $X$, the above function is analytic. A better approach might be to use a parametric representation\ \cite{Schofield}, which we have not yet tried. Equation (\ref{int_aval3a}) can be compared with the curve obtained by numerically integrating the scaling function ${\bar {\cal D}}_{-}^{(int)}$ in equation (\ref{int_aval3b}). Figure\ \ref{mf_aval_fit} shows the fit in black (equation (\ref{int_aval3a})) to the collapsed data, for curves (dashed lines) of different disorder, and system size $S=10^6$ and $S=10^7$ spins. The grey curve is the ``real'' scaling function obtained from the numerical integration of equation (\ref{int_aval3b}). Notice that the scaling collapse (done with the mean field values of the exponents $\tau+\sigma\beta\delta$ and $\sigma$) of even a system of $10^7$ spins and within $7\%$ of $R_c$ (ie. $R=0.854$) (this is the curve with the smallest peak in the graph) is not close to the ``real'' scaling function (the thick grey curve). The error is within $5\%$ for this curve (within $10\%$ for the fit). However, as $R \rightarrow R_c$, the avalanche size distribution curves seem to be approaching the ``real'' scaling function (grey curve). It is important to keep in mind when analyzing experimental or numerical data as we will in $5$ and lower dimensions, that the scaling collapse most likely does not give the limiting curve one would obtain for $1/L \rightarrow 0$ and $R \rightarrow R_c$, even for what seems like a large size, and close to the critical disorder. \begin{figure} \centerline{ \psfig{figure=Figures/Span_aval_paper_collapse_MF_legend_THESIS.ps,width=3truein} } \caption[Spanning avalanches in mean field] {{\bf Number of mean field spanning avalanches} $N_{mf}$ as a function of the disorder $R$. Curves at sizes $125$ and $343$ are not plotted. All the error bars are not shown for clarity. The ones that are shown are representative for the peaks. The error bars are smaller for larger disorders. About 26 points are used for each curve; each point being an average between $250$ (for size $512000$) and $2500$ (for size $1000$) random field configurations. The inset shows the collapse of the three largest size curves using the mean field (calculated) exponents ${\tilde \theta}= 3/8$ and $1/{\tilde \nu}=1/4$. \label{span_aval_mfa}} \end{figure} The avalanches in the avalanche size distribution are finite, by which we mean that they don't span the system. We have mentioned earlier that due to the finite size of a system, close to the critical disorder $R_c$, the largest avalanche or avalanches will span the system from one side to another. We will talk about spanning avalanches in more details later, but for now we just need to know that the number $N$ of spanning avalanches scales as $N(L,R) \sim L^\theta\ {\cal N}_{\pm}(L^{1/\nu} |r|)$ where ${\cal N}_{\pm}$ is a scaling function ($\pm$ indicates the sign of $r$), $L$ is the linear size of the system, $\theta$ is the exponent that arises from the existence of more than one spanning avalanche, and $\nu$ is the correlation length ($\xi$) exponent: $\xi \sim |r|^{-\nu}$. \begin{figure} \centerline{ \psfig{figure=Figures/1_over_Nu_span_aval_paper_MF_THESIS.ps,width=3truein} } \nobreak \centerline{ \psfig{figure=Figures/Theta_paper_MF_THESIS.ps,width=3truein} } \caption[Mean field exponents $1/{\tilde \nu}$ and ${\tilde \theta}$ from the spanning avalanches] {(a) and (b) {\bf $1/{\tilde \nu}$ and $\tilde \theta$ respectively, extracted from the mean field spanning avalanches collapses}, as a function of the geometric average of $1/S_{mf}$ for three curves collapsed together (see text). The extrapolation (non-linear for $\tilde \theta$) to $1/S_{mf} \rightarrow 0$ agrees with the calculated values for the two exponents. \label{span_aval_mf}} \end{figure} As was mentioned earlier, in mean field there is no meaning to distance or lattice, and thus there are no ``sides''. Purely for the purpose of testing our extrapolation method for finite size scaling collapses in the mean field simulation, we have defined a mean field ``spanning avalanche'' to be one with more than $\sqrt {S_{mf}}$ spins flipping at a field $H$, where $S_{mf}$ is the total number of spins in the system. (Note that the mean field exponents are valid for dimensions $6$ and above, but that in those dimensions distances do have a meaning.) Using the above definition of a mean field spanning avalanche, it can be shown (see appendix C) that the scaling form for their number is: \begin{equation} N_{mf}(S_{mf},R) \sim S_{mf}^{\tilde \theta}\ {\cal N}_{\pm}^{mf} (S_{mf}^{1/{\tilde \nu}} |r|) \label{span_aval_mf_eqn1} \end{equation} and that the values of the exponents ${\tilde \theta}$ and $1/{\tilde \nu}$ are $3/8$ and $1/4$ respectively. $N_{mf}$ is the number of mean field spanning avalanches, while ${\cal N}_{\pm}^{mf}$ is a universal scaling function. The exponents ${\tilde \theta}$ and $1/{\tilde \nu}$ are defined by the arbitrary definition for a spanning avalanche. Because of how they are defined, their values are different from the mean field values of $1/\nu = 2$ and $\theta=1$, obtained from the renormalization group\cite{Dahmen1,Dahmen2} and the exponent scaling relation $1/\sigma=(d-\theta)\nu-\beta$\cite{Dahmen1,Dahmen3}. Figure\ \ref{span_aval_mfa} shows the number of mean field spanning avalanches as a function of disorder, for several sizes, as well as the scaling collapse of the data. Note that the number of spanning avalanches close to the critical disorder $R_c=\sqrt {2/\pi}$ increases with the size $S_{mf}$ of the system, and that the peaks are getting narrower. The scaling collapse in the inset, shows only the three largest curves. For smaller sizes, the peaks do not collapse well with the larger size systems presumably due to finite size effects. The extrapolation plots for $\tilde \theta$ and $1/{\tilde \nu}$ are shown in figures\ \ref{span_aval_mf}a and \ref{span_aval_mf}b. On the horizontal axis of these two plots is the geometric mean of $1/S_{mf}$ for the three curves that are collapsed together, analogous to the extrapolation method used for the integrated avalanche size distribution. Note that the extrapolation to $1/S_{mf} \rightarrow 0$ for $\tilde \theta$ does not seem to be linear, and that the value of $1/{\tilde \nu}$ from the linear extrapolation of the $r=(R_c-R)/R$ data agrees better with the mean field value than the value obtained from the linear extrapolation of the $r=(R_c-R)/R_c$ data. Note that we measure the avalanche size distribution only for disorders at which there are no ``mean field spanning avalanches'' (for a $10^6$ system, that is for $R \ge 0.912$), since that's what we do in dimensions $2$ through $5$ (finite dimensions) to avoid large finite size effects. For the second moments of the avalanche size distribution measurements (see below), the spanning avalanches were removed (same as in finite dimensions). We have also measured the change in the magnetization $\Delta M$ due to all the spanning avalanches, as a function of the disorder $R$ (figure\ \ref{deltaM_mofhfiga}a). This gives us an independent measurement of the exponent $\beta$. In the thermodynamic limit above the critical disorder, there are no spanning avalanches so the change in the magnetization $\Delta M$ will be zero, while for small disorders the change in the magnetization will converge to one. Close and below the critical disorder $R_c$, at the critical field, the scaling form for the change in the magnetization due to the spanning avalanches will be (from equation\ (\ref{mf_mofh_equ1})): \begin{equation} \Delta M (H=H_c,R) \sim |r|^\beta. \label{deltaM_mofheq1} \end{equation} For finite size systems, as shown in the figure, the change in the magnetization is not zero above the critical disorder: the data has to be analyzed using finite size scaling. \begin{figure} \centerline{ \psfig{figure=Figures/DeltaM_mf_paper_THESIS.ps,width=3truein} } \nobreak \centerline{ \psfig{figure=Figures/DeltaM_mf_collapse_paper_THESIS.ps,width=3truein} } \caption[Magnetization change due to spanning avalanches in mean field] {(a) {\bf Change in the magnetization} due to spanning avalanches as a function of disorder $R$. The data is for several mean field system sizes. The critical disorder is $R_c =0.79788456$. The statistical errors are not larger than $0.005$ (in units of $\Delta M$). (b) Mean field scaling collapse of the change in the magnetization curves for sizes $S_{mf}=1000, 8000, 64000, 512000$. The exponents are $1/{\tilde \nu} = 0.25$ and $\beta/{\tilde \nu}= 0.125$ and $r=(R_c-R)/R$. The part of the curve that is collapsed is for $R>R_c$. \label{deltaM_mofhfiga}} \end{figure} {\noindent The dependence on the system size $S_{mf}$ can be brought in through a scaling function (see references\ \cite{Goldenfeld,Barber}) that we call $\Delta {\cal M}_{\pm}$:} \begin{equation} \Delta M(S_{mf},R) \sim |r|^\beta\ \Delta {\cal M}_{\pm}(S_{mf}^{1/{\tilde \nu}} |r|) \label{deltaM_mofheq2} \end{equation} where $\tilde \nu$ is defined above, and $\pm$ refers to the sign of $r$. We are free to define the scaling function $\Delta {\cal M}_{\pm}$ as: \begin{equation} \Delta {\cal M}_{\pm}(S_{mf}^{1/{\tilde \nu}} |r|) \equiv\ \Bigl(S_{mf}^{1/{\tilde \nu}} |r|\Bigr)^{-\beta}\ {\Delta \widetilde {\cal M}_{\pm}} (S_{mf}^{1/{\tilde \nu}} |r|), \label{deltaM_mofheq3} \end{equation} where ${\Delta \widetilde {\cal M}_{\pm}}$ is now a different scaling function. The scaling form for the change of the magnetization $\Delta M$ then becomes: \begin{equation} \Delta M(S_{mf},R) \sim S_{mf}^{-\beta/{\tilde \nu}}\ {\Delta \widetilde {\cal M}_{\pm}} (S_{mf}^{1/{\tilde \nu}} |r|). \label{deltaM_mofheq4} \end{equation} Figure\ \ref{deltaM_mofhfiga}b shows a collapse of the data using this scaling form. The collapse is done for disorders close to and above the critical disorder, that is, for $r<0$. The scaling function in figure\ \ref{deltaM_mofhfiga}b, in the range of the collapse, is therefore ${\Delta \widetilde {\cal M}_{-}}$. \begin{figure} \centerline{ \psfig{figure=Figures/1_over_Nu_deltaM_paper_MF_THESIS.ps,width=3truein} } \nobreak \centerline{ \psfig{figure=Figures/Beta_over_Nu_deltaM_paper_MF_THESIS.ps,width=3truein} } \caption[Mean field exponents $1/{\tilde \nu}$ and $\beta/{\tilde \nu}$ from the magnetization change due to spanning avalanches] {(a) and (b) {\bf Mean field exponents $1/{\tilde \nu}$ and $\beta/{\tilde \nu}$ respectively,} from collapses of the magnetization change due to spanning avalanches (see text). The extrapolation to $(1/S_{mf})_{gm} = 0$ agrees with the calculated values. \label{deltaM_mofhfigb}} \end{figure} Values for the exponents $1/{\tilde \nu}$ and $\beta/{\tilde \nu}$ extracted from such collapses at several {\it geometric average} reciprocal sizes are shown in figures\ \ref{deltaM_mofhfigb}a and \ref{deltaM_mofhfigb}b. (These plots are done the same way as for the spanning avalanches exponents.) The linear extrapolation to $1/S_{mf} = 0$ is in very good agreement with the calculated values. Note that the extrapolation for $1/\tilde \nu$ of the $r=(R_c-R)/R$ data gives again a better agreement with the calculated value than the extrapolation using the $r=(R_c-R)/R_c$ data. The exponent $\beta$ in $3$, $4$, and $5$ dimensions is calculated from $\beta/\nu$, which is extracted from the above kind of collapse. The obtained value is used to check the collapse of the $M(H)$ and $dM/dH$ data curves. \begin{figure} \centerline{ \psfig{figure=Figures/Non_span_s2_paper_and_collapse_MF_THESIS.ps,width=3truein} } \caption[Second moments of the avalanche size distribution in mean field] {{\bf Mean field second moments} of the avalanche size distribution integrated over the field $H$, for several different sizes. More than $20$ points are used for each curve; each point being an average of a few to several hundred random field configurations. The error bars for the $S_{mf}=1,000$ curve are too small to be shown. Curves at $S_{mf}=125$ and $343$ are not shown. The inset shows the collapse of these four curves at $\iota = -(\tau +\sigma\beta\delta - 3)/\sigma{\tilde \nu} = 3/8$ and $1/{\tilde \nu} = 1/4$, which are the mean field calculated values. \label{s2_mf_figa}} \end{figure} Another quantity that is related to the avalanches is the moment of the size distribution. We have measured the second, third, and fourth moment, and we will show how the second moment scales and collapses in mean field. The second moment is defined as: \begin{equation} \langle S^2 \rangle = \int S^2\ D(S,R,H,S_{mf})\ dS \label{s2_mf1} \end{equation} where $D(S,R,H,S_{mf})$ is the avalanche size distribution mentioned above, but with the system size $S_{mf}$ included as a variable since we are looking for the finite size scaling form, as is clear from the data in figure\ \ref{s2_mf_figa}. Recall that only non-spanning avalanches are included in the distribution function $D(S,R,H,S_{mf})$. Equation\ \ref{s2_mf1} can be written in terms of the scaling form for large sizes $S$ of the avalanche size distribution $D$: \begin{equation} \langle S^2 \rangle \sim \int S^2\ S^{-\tau}\ {\bar {\cal D}_\pm} (S^\sigma |r|, h/|r|^{\beta\delta}, S_{mf}^{1/\tilde \nu} |r|)\ dS \label{s2_mf2} \end{equation} As we have seen before, the dependence on the system size in the scaling function ${\bar {\cal D}_{\pm}}$ is given by $S_{mf}^{1/\tilde \nu} |r|$ where $\tilde \nu$ is defined above through the definition of a mean field spanning avalanche. If we define \begin{eqnarray} {\bar {\cal D}_\pm} (S^\sigma |r|,h/|r|^{\beta\delta}, S_{mf}^{1/\tilde \nu} |r|)\ =\ \nonumber \\ (S^\sigma |r|)^{-(2-\tau) \over \sigma}\ {\widetilde {\cal D}}_\pm (S^\sigma |r|, h/|r|^{\beta\delta}, S_{mf}^{1/\tilde \nu} |r|) \label{s2_mf2a} \end{eqnarray} where ${\widetilde {\cal D}}_\pm$ is a different scaling function, and let $u = S|r|^{1/\sigma}$, we obtain: \begin{equation} \langle S^2 \rangle\ \sim\ |r|^{(\tau-3)/\sigma} \int {\widetilde {\cal D}}_\pm (u^{\sigma}, h/|r|^{\beta\delta}, S_{mf}^{1/\tilde \nu} |r|)\ du\ \label{s2_mf3} \end{equation} The integral in equation (\ref{s2_mf3}) is a function of $h/|r|^{\beta\delta}$ and $S_{mf}^{1/\tilde \nu} |r|$ only, so we can write: \begin{equation} \langle S^2 \rangle\ \sim |r|^{(\tau-3)/\sigma}\ {\cal S}_{\pm}^{(2)}( h/|r|^{\beta\delta}, S_{mf}^{1/\tilde \nu} |r|) \label{s2_mf3a} \end{equation} which is the second moment scaling form, and ${\cal S}_{\pm}^{(2)}$ is a universal scaling function. \begin{figure} \centerline{ \psfig{figure=Figures/1_over_nu_paper_MF_non_span_s2_THESIS.ps,width=3truein} } \nobreak \centerline{ \psfig{figure=Figures/Tau_sigma_beta_delta_nu_paper_MF_THESIS.ps,width=3truein} } \caption[Mean field exponents $1/{\tilde \nu}$ and $(\tau +\sigma\beta\delta - 3)/\sigma{\tilde \nu}$ from the second moments of the avalanche size distribution] {(a) and (b) {\bf Values for $1/{\tilde \nu}$ and $(\tau +\sigma\beta\delta - 3)/\sigma{\tilde \nu}$ respectively,} extracted from the collapses of the second moments of the avalanche size distribution. The exponents are plotted as a function of the geometric average of $1/S_{mf}$ for three curves collapsed at a time (see text). The extrapolation to large sizes agrees with the calculated values for these exponents. \label{s2_mf_fig}} \end{figure} In the simulation, we have measured the second moment of the distribution integrated over the field $H$, whose scaling form can be obtained by integrating the result of equation\ (\ref{s2_mf3a}): \begin{equation} \langle S^2 {\rangle}_{\it int} \sim |r|^{(\tau-3)/\sigma} \int {\cal S}_{\pm}^{(2)}(h/|r|^{\beta\delta}, S_{mf}^{1/\tilde \nu} |r|)\ dh \label{s2_mf4} \end{equation} As was done previously, we define $u=h/|r|^{\beta\delta}$, and call the remaining integral: \begin{eqnarray} \int {\cal S}_{\pm}^{(2)}(h/|r|^{\beta\delta}, S_{mf}^{1/\tilde \nu} |r|)\ dh\ =\ \nonumber \\ (S_{mf}^{1/\tilde \nu} |r|)^{-(\tau + \sigma\beta\delta-3)/\sigma}\ {\widetilde {\cal S}}_{\pm}^{(2)}(S_{mf}^{1/\tilde \nu} |r|) \label{s2_mf4a} \end{eqnarray} to obtain the second moment of the avalanche size distribution integrated over the magnetic field $H$: \begin{equation} \langle S^2 {\rangle}_{\it int} \sim S_{mf}^{-(\tau + \sigma\beta\delta-3)/{\sigma{\tilde \nu}}}\ {\widetilde {\cal S}}_{\pm}^{(2)}(S_{mf}^{1/\tilde \nu} |r|) \label{s2_mf5} \end{equation} where ${\widetilde {\cal S}}_{\pm}^{(2)}$ is a universal scaling function ($\pm$ indicates the sign of $r$). The mean field value for $-(\tau + \sigma\beta\delta-3)/{\sigma{\tilde \nu}}$ is $3/8$. Figure\ \ref{s2_mf_figa} shows the integrated second moments of non-spanning mean field avalanches for several system sizes, and a collapse using the scaling form in equation (\ref{s2_mf5}). Figures \ref{s2_mf_fig}a and \ref{s2_mf_fig}b show the values for $1/{\tilde \nu}$ and $-(\tau + \sigma\beta\delta-3)/{\sigma{\tilde \nu}}$ respectively, for several {\it geometric average} reciprocal sizes, and show how well they linearly extrapolate to $1/S_{mf} \rightarrow 0$. These plots are done the same way as for the mean field spanning avalanches. Notice that for $1/\tilde \nu$, the linear extrapolation of the data using $r=(R_c-R)/R$ gives a much better agreement with the calculated value than the linear extrapolation of the data obtained using $r=(R_c-R)/R_c$. To summarize this section, we have shown that the values of the critical exponents extracted from our mean field simulation by scaling collapses, extrapolate to the expected (calculated) values for $R \rightarrow R_c$ and $1/S_{mf} \rightarrow 0$. Thus corrections to scaling due to finite sizes as well as finite size effects near the critical point seem to be avoided by extrapolation. The same extrapolation method is therefore used for extracting exponents in $3$, $4$, and $5$ dimensions, which we will see next. The results in $2$ dimensions will be shown last. \section{The Model} To model the long-range, far from equilibrium, collective behavior mentioned in the previous section, we define\cite{Sethna} spins $s_i$ on a hypercubic lattice, which can take two values: $s_i = \pm 1$. The spins interact ferromagnetically with their nearest neighbors with a strength $J_{ij}$, and are sitting in a uniform magnetic field $H$ (which is directed along the spins). Dirt is simulated by a random field $h_i$, associated with each site of the lattice, which is given by a gaussian distribution function $\rho (h_i)$: \begin{equation} \rho (h_i) = {1 \over {\sqrt {2\pi}} R}\ e^{-{h_i}^2 \over 2R^2} \label{model_equ1} \end{equation} of width proportional to $R$ which we call the disorder parameter, or just disorder. The hamiltonian is then \begin{equation} {\cal H} = - \sum_{<i,j>} J_{ij} s_i s_j - \sum_{i} (H + h_i) s_i \label{model_equ2} \end{equation} For the analytic calculation, as well as the simulation, we have set the interaction between the spins to be independent of the spins and equal to one for nearest neighbors, $J_{ij}=J=1$, and zero otherwise. The dynamics is deterministic, and is defined such that a spin $s_i$ will flip only when its local effective field $h^{ef\!f}_i$: \begin{equation} h^{ef\!f}_i = J \sum_{j} s_j + H + h_i \label{model_equ3} \end{equation} changes sign. All the spins start pointing down ($s_i=-1$ for all $i$). As the field is adiabatically increased, a spin will flip. Due to the nearest neighbor interaction, a flipped spin will push a neighbor to flip, which in turn might push another neighbor, and so on, thereby generating an avalanche of spin flips. During each avalanche, the external field is kept constant. For large disorders, the distribution of random fields is wide, and spins will tend to flip independently of each other. Only small avalanches will exist, and the magnetization curve will be smooth. On the other hand, a small disorder implies a narrow random field distribution which allows larger avalanches to occur. As the disorder is lowered, at the disorder $R=R_c$ and field $H=H_c$, an infinite avalanche in the thermodynamic system will occur for the first time, and the magnetization curve will show a discontinuity. Near $R_c$ and $H_c$, we find critical scaling behavior and avalanches of all sizes. Therefore, the system has two tunable parameters: the external field $H$ and the disorder $R$. We found from the mean field calculation\ \cite{Dahmen1,Dahmen2} and the simulation that a discontinuity in the magnetization exists for disorders $R \le R_c$, at the field $H_c(R) \ge H_c(R_c)$, but that only at $(R_c, H_c)$, do we have critical behavior. For finite size systems of length $L$, the transition occurs at the disorder $R_c^{ef\!f}(L)$ near which avalanches first begin to span the system in one of the {\it d} dimensions (spanning avalanches). The effective critical disorder $R_c^{ef\!f}(L)$ is larger than $R_c$, and $R_c^{ef\!f}(L) \rightarrow R_c$ as $L \rightarrow \infty$. \subsection{Simulation Results in $3$, $4$, and $5$ Dimensions} \subsubsection{Magnetization Curves} The magnetization as a function of the external field $H$ is measured for different values of the disorder $R$. Initially all the spins are pointing down ($s_i = -1$ for all $i$). The field is then slowly raised from a large negative value, until a spin flips. When the first spin has flipped, the external field is held constant while all the spins in the avalanche are flipping. The change in the magnetization due to this avalanche is just twice the size of the avalanche. Figure \ref{3d_MofH_fig}a shows the magnetization curves obtained from our simulation in $3$ dimensions for several values of the disorder $R$. Similar plots can be obtained in $4$ and $5$ dimensions\cite{mosaic}. As the disorder $R$ is decreased, a discontinuity (``jump'') in the magnetization curve appears. The critical disorder $R_c$ is the value of the disorder at which this discontinuity appears for the first time as the amount of disorder is decreased, for a system in the thermodynamic limit. For finite size systems, like the ones we use in our simulation, the ``jump'' will occur earlier. The effective critical disorder for a system of size $L$ is larger than the critical disorder $R_c$ ($1/L =0$). The critical disorder $R_c$ is found from finite size scaling collapses of the spanning avalanches and second moments of the avalanche size distribution which will be covered later. The values are listed in Table\ \ref{RH_table}. We have seen in mean field that the magnetization curves near the transition scale as \begin{equation} m(H,R) \sim |r|^\beta\ {\cal M}_{\pm}(h/|r|^{\beta\delta}) \label{mofh_3d_eq1} \end{equation} where $m=M(H,R)-M_c(H_c,R_c)$, $h=H-H_c$, and ${\cal M}_{\pm}$ is the corresponding scaling function. The critical magnetization $M_c$ and critical field $H_c$ are not universal quantities: in our mean field simulation and the hard--spin mean field model for our system\ \cite{Dahmen1}, both are zero; however they are non--zero quantities in a soft--spin model\ \cite{Dahmen1}. In general, the scaling variables in\ (\ref{mofh_3d_eq1}) need not be $r$ and $h$, but can instead be some ``rotated'' variables $r^\prime$ and $h^\prime$\ \cite{SouthAfrica} which to first approximation can be written as: \begin{equation} r^\prime = r +a h \label{mofh_3d_eq2} \end{equation} and: \begin{equation} h^\prime = h + br \label{mofh_3d_eq3} \end{equation} (See appendix A for these and other corrections.) The constants $a$ and $b$ are not universal and the critical exponents do not depend on them (for the mean field data $a=b=0$). In equation\ (\ref{mofh_3d_eq1}), the scaling variables $r$ and $h$ should be replaced by the ``rotated'' variables $r^\prime$ and $h^\prime$, but since the measurements in our simulation are in terms of $r$ and $h$, we rewrite the scaling form in terms of those. We find that in the leading order of scaling behavior, the magnetization scales like: \begin{equation} M(H,R)-M_c\ \sim\ |r|^{\beta}\ {\widetilde {\cal M}}_{\pm}\Bigl((h+br)/|r|^{\beta\delta}\Bigr). \label{mofh_3d_eq4} \end{equation} The correction $b\,r$ is dominant for $R \rightarrow R_c$, and can not be ignored. The opposite is true for $a\,h$ (see appendix A). From the previous equation, the parameters that need to be fitted are $M_c$, $H_c$, $\beta$, $\beta\delta$, and the ``tilting'' constant $b$. These should be found by collapsing the magnetization curves onto each other. As in mean field, we find that collapses of magnetization curves in $3$, $4$, and $5$ dimensions do not define well the value of the critical magnetization $M_c$. Furthermore, we observe strong correlations between the parameters, which lead to weak constraints on their values. \begin{figure} \centerline{ \psfig{figure=Figures/MofH_3d_L320_paper_THESIS.ps,width=3truein} } \nobreak \centerline{ \psfig{figure=Figures/MofH_3d_L320_collapse_paper_THESIS.ps,width=3truein} } \caption[Magnetization curves in $3$ dimensions] {(a) {\bf Magnetization curves in $3$ dimensions} for size $L=320$, and three values of disorder. The curves are averages of up to $48$ different random field configurations. Note the discontinuity in the magnetization for $R=2.20$. In finite size systems, the discontinuity in the magnetization curve occurs even for $R>R_c$ ($R_c=2.16$ in $3$ dimensions). (b) ``Tilted'' scaling collapse (see text) of the magnetization curves in $3$ dimensions for size $L=320$. The disorders range from $R=2.35$ to $R=3.20$ ($R>R_c$). The critical magnetization is chosen as $M_c=0.9$ from an analysis of the magnetization curves and is kept fixed during the collapse. The exponents are $\beta=0.036$, $\beta\delta=1.81$, and the critical field and disorder are $1.435$ and $2.16$ respectively. The ``tilting" parameter $b$ is $0.39$. \label{3d_MofH_fig}} \end{figure} To remove the dependence on the critical magnetization $M_c$, we can look at the collapse of $dM/dH$ which scales like: \begin{equation} {dM \over dH}(H,R)\ \sim\ |r|^{\beta-\beta\delta}\ {\widetilde {\cal M}}_{\pm} \Bigl((h+br)/|r|^{\beta\delta}\Bigr) \label{mofh_3d_eq5} \end{equation} Although $M_c$ does not appear in the above form, the other parameters are still not uniquely defined by the collapse. We find that we need to extract $\beta$ from the magnetization discontinuity ($\Delta M$) collapses, and $\beta\delta$ and $H_c$ from the binned avalanche size distribution collapses rather than from the magnetization curves themselves. Using the values obtained from these collapses, and the value of $R_c$, the ``tilting'' constant $b$ is then found from magnetization curve collapses (figure\ \ref{3d_MofH_fig}b). \begin{figure} \centerline{ \psfig{figure=Figures/dMdH_d3_L320_smooth10_less2_points_paper_THESIS.ps,width=3truein} } \nobreak \centerline{ \psfig{figure=Figures/dMdH_d3_L320_smooth10_less2_points_collapse_paper_THESIS.ps,width=3truein} } \caption[$dM/dH$ curves in $3$ dimensions] {(a) {\bf Derivative with respect to the field $H$ of the magnetization} $M$, for disorders $R=$ $2.35$, $2.4$, $2.45$, $2.5$, $2.6$, $2.7$, $2.85$, $3.0$, and $3.2$ (highest to lowest peak), {\bf in $3$ dimensions}. The curves are smoothed by $10$ data points before they are collapsed. (b) Scaling collapse of the data in (a) with $\beta =0.036$, $\beta\delta = 1.81$, $b=0.39$, $H_c= 1.435$, and $R_c=2.16$. While the curves are not collapsing onto a single curve, neither did they for the mean field theory curves (figure\ \protect\ref{mf_dmdh_fig}a). This is because the curves are still far from the critical disorder $R_c$. \label{dMdH_3d_fig}} \end{figure} Figure\ \ref{dMdH_3d_fig}a shows the curves for the derivative of the magnetization with respect to the field $H$, and figure\ \ref{dMdH_3d_fig}b shows the scaling collapse using the same exponent and parameter values as in figure\ \ref{3d_MofH_fig}b. The collapsed curves have disorders larger than the critical disorder: below $R_c$, the fluctuations are larger and the collapses are less reliable. Since we found that $b \neq 0$ ($b=0.39$ in $3$d), the scaling variables are indeed some $r^\prime$ and $h^\prime$, and not the variables we measure: $r$ and $h$. Therefore, the scaling functions will in general be functions of a different combination of scaling variables from the ones we used in mean field, where the scaling variables are $r$ and $h$. However, we find in appendix A that the measurements that are integrated over the external field $H$ remove the ``tilt'' parameter $b$ (other analytic corrections might still be important though). This is true for the integrated avalanche size distribution, the avalanche correlation (integrated over the field), the number of spanning avalanches, the moments of the avalanche size distribution, and the time distribution of avalanche sizes. In the sections that treat these measurements, we will ignore the ``rotation'' of axis to simplify the presentation. Note that the change in the magnetization $\Delta M$ due to the spanning avalanches is integrated over only a small range of external fields (wherever there are spanning avalanches). On the other hand, the binned avalanche size distribution is not integrated over the field $H$, and we therefore examine this measurement more carefully. \subsubsection{Moments of the Avalanche Size Distribution} The second moment of the avalanche size distribution was defined earlier (see the mean field simulation section). We found that the scaling form of the integrated over $H$ second moment is (equation\ \ref{s2_mf5}): \begin{equation} \langle S^2 {\rangle}_{\it int} \sim L^{-(\tau+\sigma\beta\delta - 3)/\sigma\nu}\ {\widetilde {\cal S}}_{\pm}^{(2)}(L^{1/\nu}|r|) \label{s2_5d} \end{equation} where $L$ is the linear size of the system, $r$ is the reduced disorder, $\widetilde {\cal S}_{\pm}^{(2)}$ is the scaling function, and $\nu$ is the correlation length exponent. The corrections are subdominant (appendix A). We can similarly define the third and fourth moment, with the exponent $-(\tau+\sigma\beta\delta - 3)/\sigma\nu$ replaced by $-(\tau+\sigma\beta\delta-4)/\sigma\nu$ and $-(\tau+\sigma\beta\delta -5)/\sigma\nu$ respectively. Figures\ \ref{s2_5d_fig}a and \ref{s2_5d_fig}b show the second moments data in $5$ dimensions for sizes $L=5, 10, 20,$ and $30$, and a collapse (again, results in $3$ and $4$ dimensions are similar and we have chosen to show the curves in $5$ dimensions for variety). The curves are normalized by the average avalanche size integrated over all fields $H$: $\int_{-\infty}^{+\infty} \int_{1}^{\infty} S\ D(S,R,H,L)\ dS\ dH$. The spanning avalanches and the infinite avalanche are not included in the calculation of the moments. The collapse does not include the $L=5$ curve because, due to finite size effects, this curve does not collapse well with the larger size curves. Table\ \ref{s2_5d_exp_table} shows the values of the exponents and $R_c$ from the collapses. The exponents for the third and fourth moment can be calculated from this table, and we find that they agree with the values obtained from their respective collapses. \section{Algorithm} There are several methods that can be used to simulate the above model. The simplest but most time and space (memory) consuming method starts by assigning a random field to each spin on the hypercubic lattice. At the beginning of the simulation, all the spins are pointing down. The external field $H$ is then increased by small increments, starting from a large negative value. After each increase of the field, all the spins are checked to find if one of them should flip (a spin flips when its effective field changes sign). If a spin flips, its neighbors are checked, and so on until no spins are left that can flip. Then, the external field is further increased, and the process repeated. Since the external magnetic field is increased in equal increments, a large amount of time is spent searching the lattice for spins that can flip. The increments have to be big enough to avoid searching the lattice when there are no spins that can flip, but small enough so that two or more spins far apart don't flip at the same field. This is the method used experimentally, but it is suited only for ``that kind of'' massively parallel computing. A variation on the above method, removes the searching through the lattice that is done even if there are no spins that can flip. It involves looking at all the spins, finding the next one that will flip and {\it then} increasing the external field so that it does. The average searching time for a flip is decreased, but is still very large. Far from the critical point, where spins will tend to flip independently of each other, the time for searching scales like $N^2$ where $N$ is the number of spins in the system. The search time can be further decreased if the random fields are initially ordered in a list. The first spin that will flip is the one on ``top'' of the list. The external field is increased until the effective field of the top spin become zero, and the spin flips. We then check its nearest neighbors, and so on, while keeping the external field constant. When no spins are left to flip, the external field needs to be increased again. The change in the external field $\Delta H$, necessary to flip the next spin, is found by looking for the spin whose random field $h_i$ satisfies: \begin{equation} h_i \ge - (H_{old}+\Delta H) - (2{n_{\uparrow}}-z)J \label{num_equ1} \end{equation} where $H_{old}$ is the field at which the previous spins have flipped, $z$ is the coordination number, and ${n_{\uparrow}}$ is the number of nearest neighbors pointing up ($s_j = +1 $) for spin $s_i$. In general, there will be a minimum of $z+1$ spins to check from the list, since ${n_{\uparrow}}$ can have the integer value between zero and $z$. The spin for which equation (\ref{num_equ1}) is satisfied for the smallest $\Delta H$, and for which the number of up neighbors is ${n_{\uparrow}}$, will flip. In general, more than $z+1$ spins will need to be checked because a spin can satisfy equation (\ref{num_equ1}) for some value of ${n_{\uparrow}}$ but might not have that number of up neighbors, or the spin might have already flipped. This algorithm decreases the searching time since not all the spins need to be checked to find the next spin that will flip. Our early simulation work \cite{Sethna,Dahmen2} used this method. In practice, about half of the time was spent for the $N\ log_2 N$ initial sorting of the list of random field numbers, where $N$ is the total number of spins in the system. The big drawback of this method (as for the ones mentioned above) is the huge amount of storage space needed to store the random fields, the positions of each spin, and the values of the spins. This becomes particularly important when larger size systems are simulated. The results in this paper use a more sophisticated algorithm which removes the need for a large storage space. It revolves around the idea that the change $\Delta H$ in the external field, between two avalanches, follows a probability distribution since the random fields $h_i$ are given by a Gaussian distribution. The increments $\Delta H$ in the external field should be chosen according to that distribution. The probability distribution itself is not known explicitly, but its integral from $0$ to some finite $\Delta H$ is. It is the probability, $P_{all}^{none}(\Delta H)$, that {\it no} spin will flip in the whole system during a field change less than $\Delta H$. It is given by: \begin{equation} P_{all}^{none}(\Delta H) = \Pi_{n_{\uparrow}}\ P_{n_{\uparrow}}^{none}(\Delta H) \label{simul_equ1} \end{equation} where the product is over $n_{\uparrow}=0,1,...,z$, and $P_{n_{\uparrow}}^{none}(\Delta H)$ is the probability for a down spin with $n_{\uparrow}$ up nearest neighbors not to flip when the external field changes by less than $\Delta H$: \begin{eqnarray} P_{n_{\uparrow}}^{none}(\Delta H)\ =\ \nonumber \\ \biggl(1 - {\int_{0}^{H_{local}(n_{\uparrow})} \rho (f)\ df - \int_{0}^{H_{local}^{new}(n_{\uparrow})} \rho (f)\ df \over P_{n_{\uparrow}}^{noflip} \Bigl(H_{local}(n_{\uparrow})\Bigr)} \biggr)^{N_{n_{\uparrow}}} \label{simul_equ2} \end{eqnarray} The function $\rho (f)$ is the random field distribution function, and $H_{local}(n_{\uparrow})$ and $H_{local}^{new}(n_{\uparrow})$ are defined respectively as: \begin{equation} H_{local}(n_{\uparrow}) = - H - (2n_{\uparrow} - z)J \label{simul_equ4} \end{equation} and \begin{equation} H_{local}^{new} (n_{\uparrow}) = -(H+\Delta H) - (2n_{\uparrow} - z)J. \label{simul_equ5} \end{equation} $P_{n_{\uparrow}}^{noflip}(H_{local})$ gives the probability that a spin with $n_{\uparrow}$ up nearest neighbors has not flipped {\it before} the field has reached the external magnetic field value $H$: \begin{equation} P_{n_{\uparrow}}^{noflip}\Bigl(H_{local}(n_{\uparrow})\Bigr) = {1 \over 2} + \int_{0}^{H_{local}(n_{\uparrow})} \rho (f)\ df, \label{simul_equ6} \end{equation} and $N_{n_{\uparrow}}$ is the number of {\it down} spins that have $n_{\uparrow}$ up neighbors. A field increment $\Delta H$ that has the required probability distribution is found by choosing a uniform random number between zero and one and solving for $\Delta H$ from equation\ (\ref{simul_equ1}), by setting the probability $P_{all}^{none}(\Delta H)$ equal to the value of the random number. Once the increment $\Delta H$ is known, we can find the next spin that will flip. We first calculate\ \cite{note1} the probability $P^{flip}(n_{\uparrow})$ for a down spin with $n_{\uparrow}$ up neighbors to flip at the new field $H + \Delta H$: \begin{equation} P^{flip}(n_{\uparrow}) = {R_{n_{\uparrow}} \over R_{tot}} \label{simul_equ61} \end{equation} where \begin{equation} R_{n_{\uparrow}} = {N_{n{\uparrow}}\ \rho \Bigl(H_{local}^{new}(n_{\uparrow})\Bigr) \over P_{n_{\uparrow}}^{noflip} \Bigl(H_{local}^{new}(n_{\uparrow})\Bigr)} \label{simul_equ62} \end{equation} is the rate at which down spins with $n_{\uparrow}$ up neighbors would flip, and $R_{tot}$ is the sum of the rates $R_{n_{\uparrow}}$ for all $n_{\uparrow}$. The spin that flips will have $k$ up neighbors, which is found by satisfying the following inequality: \begin{equation} \Sigma_{n_{\uparrow}=0}^{k}\ P^{flip}(n_{\uparrow}) > C > \Sigma_{n_{\uparrow}=0}^{k-1}\ P^{flip}(n_{\uparrow}) \label{simul_equ63} \end{equation} where the cutoff $C$ is a random number between $0$ and $1$. Once $k$ is known, a spin is then randomly picked from the list of down spins with $k$ up neighbors. After the first spin has flipped, its neighbors are checked. The probability for one of the neighbors, with ($n_{\uparrow} + 1$) up nearest neighbors, to flip at $H + \Delta H$, given that it has not yet flipped, is: \begin{equation} P_{next}(n_{\uparrow}, H+\Delta H) = 1 - {{{1 \over 2}\ +\ \int_{0}^{H_{local}^{new} (n_{\uparrow}+1)} \rho (f)\ df} \over {{1 \over 2}\ +\ \int_{0}^{H_{local}^{new}(n_{\uparrow})} \rho (f)\ df}} \label{simul_equ7} \end{equation} When all the neighbors have been checked, the size of the avalanche is stored, as well as all the other measurements. The external magnetic field $H$ is then incremented again by finding the next $\Delta H$, starting back with equation\ (\ref{simul_equ1}). The important characteristic of this method is that the random fields are not assigned to the spins at the beginning of the simulation, which for large system sizes decreases memory requirements tremendously (asymptotically, we use one bit per spin). This method has allowed us to simulate system sizes of up to $30000^2$, $1000^3$, $80^4$, and $50^5$ spins. The majority of the data analysis was performed on systems of sizes $7000^2$, $320^3$, $80^4$, and $30^5$. The SP1 and SP2 supercomputers at the Cornell Theory Center, and IBM RS6000 model 560 and J30 workstations were used for the simulation. Using this new algorithm, close to the critical disorder, one run (for a particular random field configuration) for a $320^3$ system took more than $1$ CPU hour on a SP1 node at the Cornell Theory Center, while it took close to $37$ CPU hours for a $800^3$ system on an IBM RS6000 model 560 workstation. Far above the critical disorder $R_c$, the simulation time increases substantially: $40\%$ above the critical disorder, for the $320^3$ system, the simulation time was six times longer than for the simulation at $10\%$ above $R_c$. \section{The Simulation Results} The following measurements were obtained from the simulation as a function of disorder R: \par $\bullet$ the magnetization $M(H,R)$ as a function of the \par external field $H$. \par $\bullet$ the avalanche size distribution integrated over the \par field $H$: $D_{int}(S,R)$. \par $\bullet$ the avalanche correlation function integrated over the \par field $H$: $G_{int}(x,R)$. \par $\bullet$ the number of spanning avalanches $N(L,R)$ as a \par function of the system length $L$, integrated over the \par field $H$. \par $\bullet$ the discontinuity in the magnetization $\Delta M (L,R)$ as \par a function of the system length $L$. \par $\bullet$ the second $\langle S^2 \rangle_{int}(L,R)$, third $\langle S^3 \rangle_{int}(L,R)$, and \par fourth $\langle S^4 \rangle_{int}(L,R)$ moments of the avalanche size \par distribution as a function of the system length $L$, \par integrated over the field $H$. \par {\noindent In addition, we have measured:} \par $\bullet$ the avalanche size distribution $D(S,H,R)$ as a \par function of the field $H$ and disorder $R$. \par $\bullet$ the distribution of avalanche times $D_{t}^{(int)}(S,t)$ as a \par function of the avalanche size $S$, at $R=R_c$, integrated \par over the field $H$. \par The data obtained from the simulation was used to find and describe the critical transition. It was analyzed using {\bf scaling collapses}. The mean field calculation\cite{Sethna,Dahmen1} for our model shows that near the critical point, the magnetization curve has the scaling form \begin{equation} M(H,R) - M_c(H_c,R_c) \sim |r|^\beta\ {\cal M}_{\pm}(h/|r|^{\beta\delta}) \label{model_equ4} \end{equation} where $M_c$ is the critical magnetization (the magnetization at $H_c$, for $R=R_c$), $r=(R_c-R)/R$ and $h=(H-H_c)$ are the reduced disorder and reduced field respectively, and ${\cal M}_{\pm}$ is a universal scaling function ($\pm$ refers to the sign of $r$). Both $r$ and $h$ are small. The critical exponent $\beta$ gives the scaling for the magnetization at the critical field $H_c$ ($h=0$). Its mean field value is $1/2$, and the mean field value of $\beta\delta$ is $3/2$. (Appendix A gives a short review on why scaling and scaling functions occur near a critical point, and why they have the form they do). The significance of scaling for experimental and numerical data is as follows\cite{Goldenfeld}. If the magnetization data, for example, is plotted against the field $H$, there would be one data curve for each disorder $R$ (fig.\ \ref{mf_mofhfig}a). While if we plot $|r|^{-\beta} M(H,R)$ against $h/|r|^{\beta\delta}$, all the curves close to $R_c$ and $H_c$ will {\bf collapse} (fig. \ref{mf_mofhfig}b) onto either one of two curves: one for $r<0$ (${\cal M}_{-}$), and one for $r>0$ (${\cal M}_{+}$). The functions ${\cal M}_{\pm}$ depend only on the combination $h/|r|^{\beta\delta}$ and not on the field $H$ and disorder $R$ separately, and are therefore {\it universal}. Usually, the exponents are unknown and scaling or data collapses are used to obtain them: the exponents are varied until all the curves lie on top of each other. This method is useful for analyzing numerical as well as experimental data, and is often preferred to ``data fitting'', as we will show. Numerical simulations and experiments are done on finite size systems. Often the properties of the system will depend on the linear size $L$. Functions that depend on the system's length are analyzed using {\bf finite size collapses}\cite{Goldenfeld,Barber}. An example is the number $N$ of spanning avalanches: $N(L,R) \sim L^{\theta}\ {\cal N}(L^{1/\nu} |r|)$ (to be explained later). If $N$ is plotted against $R$, there would be one data curve for each length $L$. The exponents $\theta$ and $\nu$ are obtained by plotting $L^{-\theta}N(L,R)$ against $L^{1/\nu} |r|$ onto one curve (the collapse), and extracting the exponents. The scaling forms we use for the collapses do not include corrections that are present when the data is {\it not taken} in the limit $R \rightarrow R_c$ and $L \rightarrow \infty$ (see appendix A for corrections that exist in those limits). On the other hand, finite size effects close to $R_c$ become important. It is thus necessary to extrapolate to $R \rightarrow R_c$ and $L \rightarrow \infty$ to obtain the correct exponents. We have done a mean field simulation to test our extrapolation method. The mean field exponents can be calculated analytically\ \cite{Sethna,Dahmen1}, but it is useful to check that the numerical results from the mean field simulation, for disorders away from $R_c$ and for finite sizes, extrapolate to the analytical values at $R=R_c$ and $1/L=0$. We will see that this indeed occurs, and we will use the same extrapolation method in $3$, $4$, and $5$ dimensions. The mean field simulation was done with the same code, but with some changes. In mean field, the interactions between spins are infinite in range (each spin interacts when every spin in the system with the same interaction). This means that distances and positions are not relevant, and therefore we don't need to keep track of the spins and their neighbors; we just need to know the total number of flipped spins, and the value of the external field $H$. The following section will show the results of the mean field simulation and explain the extrapolation method. Then, we will turn to results in $3$, $4$, and $5$ dimensions. And finally, we will cover the more subtle scaling behavior in two dimensions. \section{Acknowledgments} We acknowledge the support of DOE Grant \#DE-FG02-88-ER45364 and NSF Grant \#DMR-9419506. We would like to thank Sivan Kartha and Bruce W. Roberts for their initial ideas on the "probabilities'' algorithm. Furthermore, we would like to thank M. E. J. Newman, J. A. Krumhansl, J. Souletie, and M. O. Robbins for helpful conversations. This work was conducted on the SP1 and SP2 at the Cornell National Supercomputing Facility (CNSF), funded in part by the National Science Foundation, by New York State, and by IBM, and on IBM 560 workstations and the IBM J30 SMP system (both donated by IBM). We would like to thank CNSF and IBM for their support. Further pedagogical information using Mosaic is available at http://www.lassp.cornell.edu/sethna/hysteresis. \input Appendices \subsubsection{Spanning Avalanches} The critical disorder $R_c$ was defined earlier as the disorder $R$ at which an ${\it infinite}$ avalanche first appears in the system, in the thermodynamic limit, as the disorder is lowered. At that point, the magnetization curve will show a discontinuity at the magnetization $M_c(R_c)$ and field $H_c(R_c)$. For each disorder $R$ below the critical disorder, there is ${\it one}$ infinite avalanche that occurs at a critical field $H_c(R)>H_c(R_c)$\ \cite{Dahmen1,Dahmen2}, while above $R_c$ there are only finite avalanches. This is the behavior for an infinite size system. In a finite size system far below and above $R_c$ the above picture is still true, but close to the critical disorder, as we approach the transition, the avalanches get larger and larger, and we expect that one of them will be on the order of the system size and span the system from one ``side'' to another in at least one direction. This avalanche is not the infinite avalanche; it is only the largest avalanche that occurs close to the critical point. If the system was larger, this avalanche would be non--system spanning. Such an avalanche (which spans the system) we call a spanning avalanche. In our numerical simulation, we find that for finite sizes $L$, there are not one but ${\it many}$ such avalanches in $4$ and $5$ dimensions (and maybe $3$), and that their number increases as the system size increases. Figures\ \ref{span_aval_345fig}(a-c) show the number of spanning avalanches as a function of disorder $R$, for different sizes and dimensions. In $4$ and $5$ dimensions, the spanning avalanche curves become more narrow as the system size is increased. Also, the peaks shift toward the critical value of the disorder ($4.1$ and $5.96$ respectively), and the number of spanning avalanches at $R_c$ increases. This suggests that in $4$ and $5$ dimensions, for $L \rightarrow \infty$, there will be one infinite avalanche below $R_c$, none above, and an infinite number of spanning avalanches at the critical disorder $R_c$. (These spanning avalanches are infinite avalanches for $L \rightarrow \infty$.) In $3$ dimensions, the results are not conclusive, which can be noticed from figure\ \ref{span_aval_345fig}a, but also from the value of the spanning avalanche exponent $\theta = 0.15 \pm 0.15$ defined below (a value of $0$ implies only one infinite or spanning avalanche at $R_c$ as $L \rightarrow \infty$). \begin{figure} \centerline{ \psfig{figure=Figures/Span_aval_paper_d3_THESIS.ps,width=3truein} } \nobreak \centerline{ \psfig{figure=Figures/Span_aval_paper_d4_legend_THESIS.ps,width=3truein} } \nobreak \centerline{ \psfig{figure=Figures/Span_aval_paper_d5_legend_THESIS.ps,width=3truein} } \nobreak \centerline{ \psfig{figure=Figures/Span_aval_d4_collapse_paper_THESIS.ps,width=3truein} } \caption[Spanning avalanches in $3$, $4$, and $5$ dimensions] {(a) {\bf Number of spanning avalanches $N$ in $3$ dimensions,} occurring in the system between $H= -\infty$ to $H=\infty$, as a function of the disorder $R$, for linear sizes $L$: $20$ (dot-dashed), $40$ (long dashed), $80$ (dashed), $160$ (dotted), and $320$ (solid). The critical disorder $R_c$ is at $2.16$. The error bars for each curve tend to be smaller than the peak error bar for disorders above the peak and larger for disorders below the peak. They are not given here for clarity. Note that the number of avalanches increases only slightly as the size is increased. (b) {\bf Number of spanning avalanches in $4$ dimensions.} The critical disorder is $4.1$. (c) {\bf Number of spanning avalanches in $5$ dimensions.} The critical disorder is $5.96$. Both in $4$ and $5$ dimensions, the peaks grow and shift towards $R_c$ as the size of the system is increased. (d) Collapse of the spanning avalanche curves in $4$ dimensions for linear sizes $L=20,40$, and $80$. The exponents are $\theta = 0.32$ and $\nu = 0.89$, and the critical disorder is $R_c = 4.10$. The collapse is done using $r = (R_c-R)/R$. \label{span_aval_345fig}} \end{figure} In percolation, a similar multiplicity of infinite clusters\ \cite{Arcangelis,Stauffer} (as the system size is increased) is found for dimensions above $6$ which is the upper critical dimension (UCD). The UCD is the dimension at and above which the mean field exponents are valid. Below $6$ dimensions, there is only one such infinite cluster. The existence of a diverging number of infinite clusters in percolation is associated with the breakdown of the hyperscaling relation above $6$ dimensions. Since a hyperscaling relation is a relation between critical exponents that includes the dimension $d$ of the system, it is always only satisfied up to and including the upper critical dimension. In our system, the upper critical dimension is also $6$, but we find spanning avalanches in dimensions even below that. In a comment by Maritan {\it {et al.}}\cite{Maritan}, it was suggested that our system should satisfy the hyperscaling relation: $d\nu-\beta = 1/\sigma$ which is also the one found in percolation\ \cite{Stauffer}. But since our system has spanning avalanches below the upper critical dimension, this hyperscaling relation breaks down below $6$ dimensions. Due to the existence of many spanning avalanches near $R_c$, the new ``violation of hyperscaling'' relation for dimensions $3$ and above becomes\ \cite{Dahmen1,Dahmen3}: \begin{equation} (d-\theta)\nu - \beta = 1/\sigma \label{span_aval_eqn1} \end{equation} where $\theta$ is the ``breakdown of hyperscaling'' or spanning avalanches exponent defined below. One can check that our exponents in $3$, $4$, and $5$ dimensions and mean field satisfy this equation (see Tables~\ref{measured_exp_table} and~ \ref{calculated_exp_table}). For the simulation, we define a spanning avalanche to be an avalanche that spans the system in one direction. We average over all the directions to obtain better statistics. Depending on the size and dimension of the system and the distance from the critical disorder, the number of spanning avalanches for a particular value of disorder $R$ is obtained by averaging over as few as $5$ to as many as $2000$ different random field configurations. We define the exponent $\theta$ such that the number $N$ of spanning avalanches, at the critical disorder $R_c$, increases with the linear system size as: $N \sim L^{\theta}$ ($\theta > 0$). The finite size scaling form\cite{Goldenfeld,Barber} for the number of spanning avalanches close to the critical disorder is: \begin{equation} N(L,R) \sim L^{\theta}\ {\cal N}_{\pm}(L^{1/\nu}|r|) \label{span_aval_eqn2} \end{equation} where $\nu$ is the correlation length exponent and ${\cal N}_{\pm}$ is the corresponding scaling function ($\pm$ indicates the sign of $r$). The corrections to scaling are subdominant as explained in appendix A. The collapse is shown in figure\ \ref{span_aval_345fig}d. The values for $\theta$ and $\nu$ from collapses of curves of sizes $L=20, 30, 40,$ and $80$ in $4$ dimensions, are shown in Table\ \ref{span_exp_4d_table}. (We show the results and collapses in $4$ dimensions here since the existence of spanning avalanches in $3$ dimensions is not conclusive.) These values are used along with the results from other collapses to obtain Table\ \ref{measured_exp_table}. In the analysis of the avalanche size distribution, magnetization, and correlation functions for $R>R_c$, how close we chose to come to the critical disorder $R_c$ was determined by the spanning avalanches: we include no values $R$ below the first value which exhibited a spanning avalanche. \subsubsection{Magnetization Discontinuity} We have mentioned earlier that in the thermodynamic limit, at and below the critical disorder $R_c$, there is a critical field $H_c(R)>H_c(R_c)$ at which the infinite avalanche occurs. Close to the critical transition, for $r$ small and $H=H_c(R)$, the change in the magnetization due to the infinite avalanche scales as (equation (\ref{mofh_3d_eq1})): \begin{equation} \Delta M(R) \sim\ r^{\beta} \label{deltaM_eq1} \end{equation} where $r=(R_c-R)/R$, while above the transition, there is no infinite avalanche. \begin{figure} \centerline{ \psfig{figure=Figures/DeltaM_d4_paper_THESIS.ps,width=3truein} } \nobreak \centerline{ \psfig{figure=Figures/DeltaM_d4_collapse_paper_THESIS.ps,width=3truein} } \caption[Magnetization change due to spanning avalanches, in $4$ dimensions] {(a) {\bf Change in the} {\bf magnetization} {\bf due to the} {\bf spanning} {\bf avalanches in} {\bf $4$ dimensions,} for several linear sizes $L$, as a function of the disorder $R$. (b) Scaling collapse of the curves in (a) using $r=(R_c-R)/R$. The exponents are $1/\nu = 1.12$ and $\beta/\nu = 0.19$, and the critical disorder is $R_c = 4.1$. \label{deltaM_4d_fig}} \end{figure} {\noindent In finite size systems, the transition is not as sharp: we have spanning avalanches above the critical disorder. If we measure the change in the magnetization due to all the spanning avalanches (the infinite avalanche is included too), the scaling form for that quantity is going to depend on the system size $L$ analogous to the scaling of the number of spanning avalanches:} \begin{equation} \Delta M(L,R) \sim\ |r|^{\beta}\ \Delta {\cal M}_{\pm}(L^{1/\nu}|r|) \label{deltaM_eq2} \end{equation} where $\Delta {\cal M}_{\pm}$ is a universal scaling function. (Since $\Delta M(L,R)$ is measured at $h^{\prime}=0$, corrections to scaling are subdominant; see also appendix A.) Defining a new universal scaling function $\Delta \widetilde {\cal M}_{\pm}$: \begin{equation} \Delta {\cal M}_{\pm}(L^{1/\nu}|r|) \equiv\ (L^{1/\nu}|r|)^{-\beta}\ \Delta \widetilde {\cal M}_{\pm}(L^{1/\nu}|r|) \label{deltaM_eq3} \end{equation} we obtain the scaling form: \begin{equation} \Delta M(L,R) \sim\ L^{-\beta/\nu}\ \Delta {\widetilde {\cal M}_{\pm}} (L^{1/\nu}|r|) \label{deltaM_eq4} \end{equation} Figures\ \ref{deltaM_4d_fig}a and \ref{deltaM_4d_fig}b show the change in the magnetization due to the spanning avalanches in $4$ dimensions, and a scaling collapse of that data (similar results exist in $3$ and $5$ dimensions). Notice that as the system size increases, the curves approach the $|r|^{\beta}$ behavior. The exponents $1/\nu$ and $\beta/\nu$ are extracted from scaling collapses (figure\ \ref{deltaM_4d_fig}b) and are listed in table\ \ref{deltaM_4d_table}. The value of $\beta$ is calculated from $\beta/\nu$ and the knowledge of $\nu$, and is the value used for collapses of the magnetization curves (see earlier). \begin{figure} \centerline{ \psfig{figure=Figures/non_span_d5_paper_THESIS.ps,width=3truein} } \nobreak \centerline{ \psfig{figure=Figures/non_span_d5_collapse_paper_THESIS.ps,width=3truein} } \caption[Second moments of the avalanche size distribution in $5$ dimensions] {(a) {\bf Second moments} of the avalanche size distribution integrated over the field $H$, {\bf in $5$ dimensions.} Error bars are largest for smaller disorders (shown on the curves). The curves have between $24$ and $50$ points, and the value of the second moment for each disorder is averaged over $3$ to $100$ different random field configurations. (b) Scaling collapse of the $L=10, 20$, and $30$ curves from (a) using $r=(R_c-R)/R$. The exponents are $1/\nu = 1.47$ and $\rho = -(\tau+\sigma\beta\delta - 3)/\sigma\nu = 2.95$, and the critical disorder is $R_c = 5.96$. \label{s2_5d_fig}} \end{figure} \subsubsection{Avalanche Time Measurement} The exponents we have measured so far are static scaling exponents: they do not depend on the dynamics of the model. If we measure the time an avalanche takes to occur, we are making a dynamical measurement. The time measurement in the numerical simulation is done by increasing the time ``meter'' by one for each shell of spins in the avalanche; it corresponds to a synchronous dynamics, where, when all unstable spins are flipped, time is incremented by one, and the new list of unstable spins is generated. The scaling relation between the time $t$ it takes an avalanche to occur and the size $S$ of that avalanche for small disorder $r$ can be found by noting that the characteristic duration of an avalanche is proportional to the correlation length $\xi$ to the power $z$\ \cite{dynamicz,MaBinney}: \begin{equation} t \sim \xi^z \label{time_equ1} \end{equation} The exponent $z$ is known as the dynamical critical exponent. Equation\ (\ref{time_equ1}) gives the scaling for the time it takes for a spin to ``feel'' the effect of another a distance $\xi$ away. Since the correlation length $\xi$ scales like $r^{-\nu}$ close to the critical disorder, and the characteristic size $S$ as $r^{-1/\sigma}$, the time $t$ then scales with large sizes as: \begin{equation} t \sim S^{\sigma \nu z} \label{time_equ2} \end{equation} \begin{figure} \centerline{ \psfig{figure=Figures/Aval_Time_3d_L800_paper_THESIS.ps,width=3truein} } \nobreak \centerline{ \psfig{figure=Figures/Aval_Time_3d_L800_paper_collapse_THESIS.ps,width=3truein} } \caption[Avalanche time distribution curves in $3$ dimensions] {(a) {\bf Avalanche time distribution curves in $3$ dimensions,} for avalanche size bins from about $2000$ to $40000$ spins (from upper left to lower right corner). The system size is $800^3$ at $R=2.26$. The curves are from only one random field configuration. (b) Scaling collapse of curves in (a). The values of the exponents are $\sigma\nu z = 0.57$ and $(\tau+\sigma\beta\delta+ \sigma\nu z)/\sigma\nu z = 4.0$. \label{time_3d_fig}} \end{figure} In our simulation, we measure the distribution of times for each avalanche size $S$. The distribution of times $D_t(S,R,H,t)$ for an avalanche of size $S$ close to the critical field $H_c$ and critical disorder $R_c$ is \begin{equation} D_t(S,R,H,t) \sim S^{-q}\ {\bar {\cal D}}_{\pm}^{(t)} (S^{\sigma}|r|, h/|r|^{\beta\delta}, t/S^{\sigma\nu z}) \label{time_equ3} \end{equation} where $q=\tau +\sigma\nu z$, and is defined such that \begin{eqnarray} \int_{-\infty}^{+\infty} \!\! \int_{1}^{\infty} D_t(S,R,H,t)\ dH\ dt\ =\ \nonumber \\ S^{-(\tau + \sigma\beta\delta)}\ {\bar {\cal D}}_{\pm}^{(int)}(S^{\sigma}|r|) \label{time_equ4} \end{eqnarray} where ${\bar {\cal D}}_{\pm}^{(int)}$ was defined in the integrated avalanche size distribution section. The avalanche time distribution integrated over the field $H$, at the critical disorder ($r=0$) is: \begin{equation} D_t^{(int)}(S,t)\ \sim\ t^{-{(\tau + \sigma\beta\delta + \sigma\nu z) /\sigma\nu z}}\ {\cal D}_t^{(int)}(t/S^{\sigma\nu z}) \label{time_equ5} \end{equation} which is obtained from equation (\ref{time_equ3}) in a derivation analogous to the one for the integrated avalanche size distribution scaling form. Figures\ \ref{time_3d_fig}a and \ref{time_3d_fig}b show the avalanche time distribution integrated over the field $H$ for different avalanche sizes, and a collapse of these curves using the above scaling form, for a $800^3$ system at $R=2.260$ (just above the range where spanning avalanches occur). The data is saved in logarithmic size bins, each about $1.2$ times larger than the previous one. The time is also measured logarithmically (next bin is $1.1$ times larger than the previous one). The extracted value for $z$ in $3$ dimensions is $1.68 \pm 0.07$. The results for other dimensions are listed in Table\ \ref{measured_exp_table}.
proofpile-arXiv_065-414
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} Extensive attention has been lavished on the overscreened multichannel Kondo model after the discover of its non-fermi liquid (NFL) behavior by Nozi\'{e}res and Blandin (NB) \cite{blandin}. NB also pointed out that lattice effects in real metals will cause the anisotropy between the two flavor channels and that in the low temperature limit, the impurity is totally screened by the strong coupling channel with the weak coupling channel unaffected. Using Numerical Renormalization Group (NRG), Ref.\cite{cox2} confirmed NB's conjecture. Using Conformal Field Theory (CFT), Ref.\cite{affleck} found a relevant dimension 1/2 operator in the flavor sector near the 2 channel Kondo (2CK) fixed point and suggested the system flows to the Fermi-liquid (FL) fixed point pointed out by NB. Using Yuval-Anderson's approach, Ref.\cite{gogolin} found a solvable line and calculated the exact crossover free energy function from the 2CK fixed point to the FL fixed point along this solvable line. It is known that in the large $ U $ limit, the ordinary one channel symmetric Anderson impurity model(AIM) can be mapped to the one channel Kondo model. However, as shown by Ref.\cite{com,coleman}, if the original $ O(4) $ symmetry of the AIM is broken to $ O(3) \times O(1) $, in the strong coupling limit, the AIM is mapped to the one channel compactified Kondo model (1CCK) where the impurity spin couples to both the spin and the isospin(charge) currents of the one channel conduction electrons. Recently, Andrei and Jerez \cite{andrei}, using Bethe Ansatz, reinvestigated the 2CFAK and conjectured that the 2CFAK flow to some new NFL fixed points. Coleman and Schofield \cite{coleman}, using strong coupling method, reinvestigated the 1CCK and claimed the system flows to another kind of non-Fermi liquid fixed point which, similar to 1-dim Luttinger liquid, has the same thermodynamics as fermi liquid but different excitation spectrum. Moreover, they claimed that the 1CCK has exactly the same low energy excitations as those of the 2CFAK, therefore concluded that their results also apply to the 2CFAK. So far, Bethe Ansatz can only calculate thermodynamic quantities of multichannel Kondo models, the correlation functions are needed to resolve if the fixed points are NFL or FL. It is important to point out that the charge degrees of freedom of the original model being removed, the 1CCK in Ref.\cite{com,coleman} has completely different transport properties, correlation functions and excitation spectrums than the original 2CFAK, although it do share the same thermodynamic properties as the 2CFAK. As emphasized by AL \cite{review}, although the boundary interactions only happen in the spin sector; the spin, flavor and charge degree of freedoms are {\em not} totally decoupled, there is a constraint( or gluing condition) to describe precisely how these degree of freedoms are combined at different boundary fixed points, the finite size spectrum is determined by this gluing condition. The boundary operator contents and the scaling dimensions of all the boundary operators are also given by the gluing condition. However, in order to find the gluing conditions at the intermediate coupling fixed points, the fusion rules should be identified which are usually difficult in Non-Abelian bosonization approach. For 4 pieces of bulk fermions, the non-interacting theory possesses $ SO(8) $ symmetry, Maldacena and Ludwig (MS) \cite{ludwig} showed that finding the gluing conditions at the fixed points are exactly equivalent to finding the boundary conditions of the fermions at the fixed points; the CFT describing the fixed points are simply free chiral bosons with the boundary conditions. In Ref.\cite{powerful}, the author developed a simple and powerful method to study certain class of quantum impurity models. The method can quickly identify all the possible boundary fixed points and their {\em maximum } symmetry, therefore avoid the difficulty of finding the fusion rules, it can also demonstrate the physical picture at the boundary explicitly. In this paper, we apply the method to study the two models. All {\em the possible} fixed points and their symmetries are identified; the finite size spectra, the electron conductivity and pairing susceptibility are calculated. All the leading and subleading irrelevant operators are identified, their corrections to the correlation functions are evaluated. In section II, {\em Taking all the degrees of freedom into account}, We show that the only NFL fixed point of the 2CFAK is the NFL fixed point of the 2CK with the symmetry $ O(3) \times O(5) $. Any flavor anisotropies between the two channels drive the system to the fermi-liquid (FL) fixed point with the symmetry $ O(4) \times O(4) $ where one of the two channels suffers the phase shift $ \pi/2 $ and the other remains free. The conventional wisdom about the 2CFAK is rigorously shown to be correct. In section III, we repeat the same program to the 1CCK. We find that the NFL fixed point of the 1CCK has the symmetry $ O(3) \times O(1) $ and has the same thermodynamics as the NFL fixed point of the 2CK. The finite size spectrum is listed and compared with that of the 2CK. However, {\em in contrast to} the 2CK, its conductivity shows $ T^{2} $ bahaviour and there is {\em no} pairing susceptibility enhancement. Any anisotropies between the spin and isospin sectors drive the system to the FL fixed point with the symmetry $ O(4) $ where the electrons suffer the phase shift $ \pi/2 $. The finite size spectrum of this FL fixed point is also listed and compared with that of the 2CFAK. In section IV, we conclude and propose some open questions. Finally, in the appendix, we study the stable FL fixed point of the 2CFAK using Non-Abelian bosonization and compare with the Abelian bosonization calculations done in section II. \section{The two channel flavor anisotropic Kondo model} The Hamiltonian of the 2CFAK is: \begin{eqnarray} H &= & i v_{F} \int^{\infty}_{-\infty} dx \psi^{\dagger}_{i \alpha }(x) \frac{d \psi_{i \alpha }(x)}{dx} + \sum_{a=x,y,z} \lambda^{a} ( J^{a}_{1}(0)+J^{a}_{2}(0) ) S^{a} + \sum_{a=x,y,z} \alpha^{a} (J^{a}_{1}(0)-J^{a}_{2}(0)) S^{a} \nonumber \\ & + & h ( \int dx J^{z}_{s}(x) + S^{z} ) \label{kondob} \end{eqnarray} where $ J^{a}_{i}(x) =\frac{1}{2} \psi^{\dagger}_{i \alpha }(x) \sigma^{a}_{\alpha \beta} \psi_{i \beta }(x) $ are the spin currents of the channel $ i=1,2 $ conduction electrons respectively. $ \alpha^{a} =0, \pm \lambda^{a} $ correspond to the 2CK and the one channel Kondo model respectively. If $ \lambda^{a} =\lambda, \alpha^{a} =\alpha \neq 0 $ , the above Hamiltonian breaks $ SU(2)_{s} \times SU(2)_{f} \times U(1)_{c} $ symmetry of the 2CK to $ SU(2)_{s} \times U(1)_{f} \times U(1)_{c} $ ( or equivalently $ SU(2)_{s} \times U(1)_{c1} \times U(1)_{c2}$, because we have two independent U(1) charge symmetries in the channel 1 and the channel 2 ). In this section, for simplicity, we take $ \lambda^{x}=\lambda^{y}=\lambda, \lambda^{z} \neq \lambda; \alpha^{x}=\alpha^{y}=\alpha, \alpha^{z} \neq \alpha $. The symmetry in the spin sector is reduced to $ U(1) \times Z_{2} \sim O(2) $ \cite{semi}. In the following, we closely follow the notations of Emery-Kivelson \cite{emery}. Abelian-bosonizing the four bulk Dirac fermions separately: \begin{equation} \psi_{i \alpha }(x )= \frac{P_{i \alpha}}{\sqrt{ 2 \pi a }} e ^{- i \Phi_{i \alpha}(x) } \label{first} \end{equation} Where $ \Phi_{i \alpha} (x) $ are the real chiral bosons satisfying the commutation relations \begin{equation} [ \Phi_{i \alpha} (x), \Phi_{j \beta} (y) ] = \delta_{i j} \delta_{\alpha \beta} i \pi sgn( x-y ) \end{equation} The cocyle factors have been chosen as: $ P_{1 \uparrow}= P_{1 \downarrow} = e^{i \pi N_{1 \uparrow} }, P_{2 \uparrow}= P_{2 \downarrow} = e^{i \pi ( N_{1 \uparrow} + N_{1 \downarrow} + N_{2 \uparrow} ) } $. It is convenient to introduce the following charge, spin, flavor, spin-flavor bosons: \begin{eqnarray} \Phi_{c} & = & \frac{1}{2} ( \Phi_{1 \uparrow }+ \Phi_{1 \downarrow }+ \Phi_{2 \uparrow }+ \Phi_{2 \downarrow } ) \nonumber \\ \Phi_{s} & = & \frac{1}{2} ( \Phi_{1 \uparrow }- \Phi_{1 \downarrow }+ \Phi_{2 \uparrow }- \Phi_{2 \downarrow } ) \nonumber \\ \Phi_{f} & = & \frac{1}{2} ( \Phi_{1 \uparrow }+ \Phi_{1 \downarrow }- \Phi_{2 \uparrow }- \Phi_{2 \downarrow } ) \nonumber \\ \Phi_{sf}& = & \frac{1}{2} ( \Phi_{1 \uparrow }- \Phi_{1 \downarrow }- \Phi_{2 \uparrow }+ \Phi_{2 \downarrow } ) \label{second} \end{eqnarray} The spin currents $ J^{a}(x) = J^{a}_{1}(x) + J^{a}_{2}(x) $ and $ \tilde{J}^{a}(x) = J^{a}_{1}(x) - J^{a}_{2}(x) $ can be expressed in terms of the above chiral bosons \begin{eqnarray} J_{x}= \frac{1}{\pi a} \cos \Phi_{s} \cos \Phi_{sf},~ J_{y}= \frac{1}{\pi a} \sin \Phi_{s} \cos \Phi_{sf},~ J_{z}= -\frac{1}{ 2 \pi} \frac{ \partial \Phi_{s}}{\partial x} \nonumber \\ \tilde{J}_{x}=- \frac{1}{\pi a} \sin \Phi_{s} \sin \Phi_{sf},~ \tilde{J}_{y}= \frac{1}{\pi a} \cos \Phi_{s} \sin \Phi_{sf},~ \tilde{J}_{z}= -\frac{1}{ 2 \pi} \frac{ \partial \Phi_{sf}}{\partial x} \label{current} \end{eqnarray} After making the canonical transformation $ U= \exp [ i S^{z} \Phi_{s}(0)] $ and the following refermionization \begin{eqnarray} S^{x} &= & \frac{ \widehat{a}}{\sqrt{2}} e^{i \pi N_{sf}},~~~ S^{y}= \frac{ \widehat{b}}{\sqrt{2}} e^{i \pi N_{sf}},~~~ S^{z}= -i \widehat{a} \widehat{b} \nonumber \\ \psi_{sf} & = & \frac{1}{\sqrt{2}}( a_{sf} - i b_{sf} ) = \frac{1}{\sqrt{ 2 \pi a}} e^{i \pi N_{sf}} e^{-i \Phi_{sf} } \nonumber \\ \psi_{s,i} & = & \frac{1}{\sqrt{2}}( a_{s,i} - i b_{s,i} )= \frac{1}{\sqrt{ 2 \pi a}} e^{i \pi( d^{\dagger}d + N_{sf})} e^{-i \Phi_{s} } \label{refer} \end{eqnarray} The transformed Hamiltonian $ H^{\prime}= U H U^{-1} = H_{sf} + H_{s} + \delta H $ can be written in terms of the Majorana fermions \cite{atten}: \begin{eqnarray} H_{sf} &= & \frac{ i v_{F} }{2} \int dx (a_{sf}(x) \frac{ \partial a_{sf}(x)} {\partial x} + b_{sf}(x) \frac{ \partial b_{sf}(x)} {\partial x} ) -i \frac{ \lambda }{\sqrt{ 2 \pi a}} \widehat{a} b_{sf}(0) +i \frac{ \alpha }{\sqrt{ 2 \pi a}} \widehat{b} a_{sf}(0) \nonumber \\ H_{s}& = & \frac{ i v_{F} }{2} \int dx (a_{s}(x) \frac{ \partial a_{s}(x)} {\partial x} + b_{s}(x) \frac{ \partial b_{s}(x)} {\partial x} ) -i h \int dx a_{s}(x) b_{s}(x) \nonumber \\ \delta H &= & -\lambda_{z}^{\prime} \widehat{a} \widehat{ b} a_{s}(0) b_{s}(0) -\alpha_{z} \widehat{a} \widehat{b} a_{sf}(0) b_{sf}(0) \label{anderson} \end{eqnarray} where $ \lambda_{z}^{\prime} = \lambda^{z} - 2 \pi v_{F} $. It is instructive to compare the above equation with Eq.3 in Ref.\cite{sf}. They looks very similar: {\em half} of the impurity spin coupled to half of the spin-flavor electrons, {\em another half} of the impurity spin coupled to {\em another} half of the spin-flavor electrons. However {\em the two canonical transformations employed in the two models are different}. This fact make the boundary conditions of this model rather different from that of the two channel spin-flavor Kondo model (2CSFK) discussed in Ref.\cite{sf}. The above Hamiltonian was first derived by Ref.\cite{gogolin} using Anderson-Yuval's approach. They found the solvable line $ \lambda^{z} =2 \pi v_{F}, \alpha^{z}=0 $ and calculated the exact crossover function of free energy along this solvable line. Using EK's method, We rederived this Hamiltonian \cite{trivial}. The huge advantage of EK's method over Anderson-Yuval's approach is that the {\em boundary conditions} at different boundary fixed points can be identified \cite{powerful}. By using the Operator Product Expansion (OPE) of the various operators in Eq.\ref{anderson} \cite{cardy}, we get the RG flow equations near the weak coupling fixed point $\lambda_{z}=2 \pi v_{F}, \lambda=\alpha=\alpha_{z}=0 $ \begin{eqnarray} \frac{ d \lambda}{d l} & = &\frac{1}{2} \lambda+ \alpha \alpha_{z} \nonumber \\ \frac{ d \alpha}{d l} & = &\frac{1}{2} \alpha - \lambda \alpha_{z} \nonumber \\ \frac{ d \alpha_{z}}{d l} & = & -\lambda \alpha \label{danger} \end{eqnarray} The fact that we find {\em two} relevant operators in the above equations may indicate there are {\em two} intermediate coupling fixed points. However, in the following, the two intermediate coupling fixed points are shown to be the same. The {\em original} impurity spin in $ H $ are related to those in $ H^{\prime} $ by \begin{eqnarray} S^{H}_{x} &= & U S_{x} U^{-1} = S_{x} \cos \Phi_{s}(0) - S_{y} \sin \Phi_{s}(0) \nonumber \\ S^{H}_{y} &= & U S_{y} U^{-1} = S_{x} \sin \Phi_{s}(0) + S_{y} \cos \Phi_{s}(0) \nonumber \\ S^{H}_{z} &= & U S_{z} U^{-1} = S_{z} \label{change} \end{eqnarray} Using the refermionization Eq.\ref{refer}, the {\em original} impurity spin in $ H $ can be written in terms of fermions \begin{eqnarray} S^{H}_{x} &= & i( \widehat{b} a_{s,i}+\widehat{a} b_{s,i} ) \nonumber \\ S^{H}_{y} &= & i( \widehat{b} b_{s,i}-\widehat{b} a_{s,i} ) \nonumber \\ S^{H}_{z} &= & -i \widehat{a} \widehat{b} \label{imp} \end{eqnarray} At $ \lambda^{\prime}_{z}=0 $, the spin boson $ \Phi_{s} $ completely decouples from the impurity in $ H^{\prime} $, therefore $ \chi_{imp} =0 $. Because the canonical transformation $ U $ is a boundary condition changing operator \cite{boundary,powerful}, at $ \lambda^{\prime}_{z} =0 $, this leads to \begin{equation} a^{s}_{L}(0)=-a^{s}_{R}(0), ~~ b^{s}_{L}(0)=-b^{s}_{R}(0) \label{bound1} \end{equation} Following Ref.\cite{powerful}, in order to identify the fixed points along the solvable line $ \lambda^{\prime}_{z}=0, \alpha_{z}=0 $ (we also set $ h=0 $), we write $ H_{sf} $ in the action form \begin{eqnarray} S &= & S_{0} + \frac{\gamma_{1}}{2} \int d \tau \widehat{a}(\tau) \frac{\partial \widehat{a}(\tau)}{\partial \tau} + \frac{\gamma_{2}}{2} \int d \tau \widehat{b}(\tau) \frac{\partial \widehat{b}(\tau)}{\partial \tau} \nonumber \\ & - & i \frac{ \lambda }{\sqrt{ 2 \pi a}} \int d \tau \widehat{a}(\tau) b_{sf}(0, \tau) +i \frac{ \alpha }{\sqrt{ 2 \pi a}} \int d \tau \widehat{b}(\tau) a_{sf}(0,\tau) \label{action} \end{eqnarray} When performing the RG analysis of the action $ S $, we keep \cite{above} 1: $ \gamma_{2}=1, \lambda $ fixed, 2: $ \gamma_{1}=1, \alpha $ fixed, 3: $ \lambda, \alpha $ fixed; three fixed points of Eq.\ref{anderson} can be identified \subsection{ Fixed point 1} This fixed point is located at $ \gamma_{1}=0, \gamma_{2}=1 $ where $ \widehat{b} $ decouples, but $ \widehat{a} $ loses its kinetic energy and becomes a Grassmann Lagrangian multiplier. Integrating $\widehat{a} $ out leads to the following boundary conditions \cite{trick}: \begin{equation} b^{sf}_{L}(0)=-b^{sf}_{R}(0) \label{bound2} \end{equation} Eqs.\ref{bound1},\ref{bound2} can be expressed in terms of bosons: \begin{equation} \Phi_{s,L}(0)=\Phi_{s,R}(0)+\pi, ~~~ \Phi_{sf,L}(0)=-\Phi_{sf,R}(0)+\pi \end{equation} This is just the non-fermi liquid fixed point of the 2CK. The three Majorana fermions in the spin sector being twisted, this fixed point possesses the symmetry $ O(3) \times O(5) $. The finite size spectrum of this fixed point was listed in Ref.\cite{powerful}. The local correlation functions at the 2CK fixed point are \cite{powerful}: \begin{equation} \langle \widehat{a}( \tau ) \widehat{a}(0) \rangle =\frac{1}{\tau},~~~~ \langle b_{sf}( \tau ) b_{sf}(0) \rangle =\frac{\gamma^{2}_{1}}{\tau^{3}} \label{dimension} \end{equation} From the above equation, we can read the scaling dimensions of the various fields $ [\widehat{b}]=0, [\widehat{a}]=[a_{s}]=[b_{s}]=[a_{sf}]=1/2, [b_{sf}]=3/2 $. As shown in Ref.\cite{powerful}, at the fixed point, the impurity degree of freedoms completely disappear: $\widehat{b} $ decouples and $ \widehat{a} $ turns into the {\em non-interacting } scaling field at the fixed point \cite{care} \begin{equation} \widehat{a} \sim b_{sf}(0,\tau) \end{equation} Using Eq.\ref{imp}, the impurity spin turns into \begin{eqnarray} S^{H}_{x}(\tau) &= & i( \widehat{b} a_{s,i}(0,\tau)+b_{sf}(0,\tau) b_{s,i}(0,\tau) ) \nonumber \\ S^{H}_{y}(\tau) &= & i( \widehat{b} b_{s,i}(0,\tau)-b_{sf}(0,\tau) a_{s,i}(0,\tau) ) \nonumber \\ S^{H}_{z}(\tau) &= & i \widehat{b} b_{sf}(0,\tau) \end{eqnarray} Using the relation \begin{equation} \psi^{H}_{s}(x)= U \psi_{s}(x) U^{-1}=i sgnx \psi_{s,i}(x) \label{reverse} \end{equation} We get \cite{cut} \begin{eqnarray} S_{x}(\tau) &= & i( -\widehat{b} b_{s}(0,\tau)+b_{sf}(0,\tau) a_{s}(0,\tau) ) \nonumber \\ S_{y}(\tau) &= & i( \widehat{b} a_{s}(0,\tau)+b_{sf}(0,\tau) b_{s}(0,\tau) ) \nonumber \\ S_{z}(\tau) &= & i ( \widehat{b} b_{sf}(0,\tau) + a_{s}(0,\tau) b_{s}(0,\tau) ) \label{add} \end{eqnarray} The impurity spin-spin correlation function $ \langle S^{a}(\tau) S^{a}(0) \rangle =\frac{1}{\tau} $. The above equations \cite{add} are consistent with the CFT identifications \cite{line} \begin{equation} \vec{S} \sim \vec{\phi} + \vec{ J} + \cdots \end{equation} The 2CK fixed point is unstable, because there is a dimension 1/2 relevant operator $ \widehat{b} a_{sf} $, the OPE of $ a_{sf} $ with itself will generate the dimension 2 energy momentum tensor of this Majorana fermion $ T(\tau)= \frac{1}{2} a_{sf}(0,\tau) \frac{ \partial a_{sf}(0,\tau)}{\partial \tau} $, The OPE of the energy momentum tensor with the primary field $ a_{sf} $ is \begin{equation} T(\tau_{1}) a_{sf}(\tau_{2})= \frac{ \frac{1}{2} a_{sf}(\tau_{2})}{ (\tau_{1}-\tau_{2})^{2}} + \frac{ L_{-1} a_{sf}(\tau_{2})}{\tau_{1}-\tau_{2}} + L_{-2} a_{sf}(\tau_{2}) + \cdots \end{equation} First order descendant field of this primary field $ L_{-1} a_{sf}(0,\tau)= \frac{ \partial a_{sf}(0,\tau)}{\partial \tau} $ with dimension 3/2 is generated. $ \lambda^{\prime}_{z} $ term in $ \delta H $ has scaling dimension 3/2, it will generate a dimension 2 operator $ a_{s}(0,\tau) \frac{ \partial a_{s}(0,\tau)}{\partial \tau} + b_{s}(0,\tau) \frac{ \partial b_{s}(0,\tau)}{\partial \tau} $. $\gamma_{2} $ term has dimension 2 also. From Eq.\ref{dimension}, we can see $ \alpha_{z} $ term has scaling dimension 5/2, it can be written as \begin{equation} :\widehat{a}(\tau) \frac{\partial \widehat{a}(\tau)}{\partial \tau}: a_{sf}(0,\tau) = :b_{sf}(0,\tau) \frac{\partial b_{sf}(0,\tau)}{\partial \tau}: a_{sf}(0,\tau) \end{equation} The bosonized form of this operator is \begin{equation} :( \cos 2\Phi_{sf}(0,\tau)-\frac{1}{2} (\partial \Phi_{sf}(0,\tau))^{2}): \sin \Phi_{sf}(0,\tau) \end{equation} Using CFT, Ref.\cite{affleck} predicted a dimension 1/2 relevant operator $ \phi^{3}_{f} $ in the flavor sector. Ref.\cite{line} classified all the first order descendants of the primary operator in the spin sector. In the flavor sector, the same classification apply, $ \vec{J}_{-1} \cdot \vec{\phi}_{f} $ is Charge-Time Reversal (CT) odd, therefore is not allowed, but $ L_{-1} \phi^{3}_{f} $ is CT even. The CFT analysis is completely consistent with the above EK's solution. In order to make this fixed point stable, we have to tune $ \alpha =\alpha_{z}=0 $, namely the channel anisotropy is strictly prohibited. If $\alpha=0 $, but $ \alpha_{z} \neq 0 $, because $ \alpha_{z} $ is highly irrelevant, it {\em seems} the 2CK fixed point is stable. However, this is not true. From the RG flow Eq.\ref{danger}, it is easy to see that even initialy $ \alpha=0 $, it will be generated, $ \alpha_{z} $ is 'dangerously' irrelevant. \subsection{ Fixed point 2} This fixed point is located at $ \gamma_{1}=1, \gamma_{2}=0 $ where $ \widehat{a} $ decouples, but $ \widehat{b} $ loses its kinetic energy and becomes a Grassmann Lagrangian multiplier. Integrating $\widehat{b} $ out leads to the following boundary conditions: \begin{equation} a^{sf}_{L}(0)=-a^{sf}_{R}(0) \label{dual} \end{equation} Eqs.\ref{bound1},\ref{dual} can be expressed in terms of bosons: \begin{equation} \Phi_{s,L}(0)=\Phi_{s,R}(0)+\pi, ~~~ \Phi_{sf,L}(0)=-\Phi_{sf,R}(0) \end{equation} This fixed point also possesses the symmetry $ O(3) \times O(5) $. In fixed points 1 and 2, $\widehat{a} $ and $\widehat{b} $, $ b_{sf} $ and $ a_{sf} $ exchange roles. As discussed in the fixed point 1, $ \alpha_{z} $ is 'dangerously' irrelevant. In order to make this fixed point stable, we have to tune $ \lambda =\alpha_{z}=0 $. This fixed point is actually {\em the same} with the 2CK fixed point. This can be seen most clearly from the original Eq.\ref{kondob}: if $\lambda=\alpha_{z} =0 $, under the $ SU(2) $ transformation on the channel 2 fermions $ \psi_{2 \uparrow} \rightarrow i \psi_{2 \uparrow}, \psi_{2 \uparrow} \rightarrow -i \psi_{2 \uparrow} $, the spin currents of channel 2 transform as $ J^{x}_{2} \rightarrow -J^{x}_{2}, J^{y}_{2} \rightarrow -J^{y}_{2}, J^{z}_{2} \rightarrow J^{z}_{2} $, Eq.\ref{kondob} is transformed back to the 2 channel flavor symmetric Kondo model. This can also be seen from Eq.\ref{current}, $ \tilde{J}_{x}, \tilde{J}_{y}, J_{z} $ also satisfy the $ \widehat{SU}_{2}(2) $ algebra. \subsection{ Fixed point 3} This fixed point is located at $ \gamma_{1}=\gamma_{2}=0 $ where both $\widehat{a} $ and $ \widehat{b} $ lose their kinetic energies and become two Grassmann Lagrangian multipliers. Integrating them out leads to the following boundary conditions: \begin{equation} b^{sf}_{L}(0)=-b^{sf}_{R}(0), ~~ a^{sf}_{L}(0)=-a^{sf}_{R}(0) \label{bound3} \end{equation} Eqs.\ref{bound1}, \ref{bound3} can be expressed in term of bosons: \begin{equation} \Phi^{s}_{L}=\Phi^{s}_{R} + \pi, ~~~ \Phi^{sf}_{L}=\Phi^{sf}_{R} + \pi \end{equation} Substituting the above equation to Eqs. \ref{first} \ref{second} and paying attention to the {\em spinor} nature of the representation \cite{hopping}, it is easy to see that depending on the sign of $\alpha$, {\em one} of the two channels suffer $\frac{\pi}{2} $ phase shift, {\em the other} remains free. The four Majorana fermions being twisted, this fixed point has the symmetry $ O(4) \times O(4) $ with $ g=1 $. The finite size spectrum of this fixed point is listed in Table \ref{flavor}, it is the sum of that with phase shift $ \pi/2 $ and that of free electrons. This scenario is completely consistent with NRG results of Ref.\cite{cox2}. The local correlation functions at the FL fixed point are \cite{powerful}: \begin{eqnarray} \langle \widehat{a}( \tau ) \widehat{a}(0) \rangle =\frac{1}{\tau},~~~~ \langle b_{sf}( \tau ) b_{sf}(0) \rangle =\frac{\gamma^{2}_{1}}{\tau^{3}} \nonumber \\ \langle \widehat{b}( \tau ) \widehat{b}(0) \rangle =\frac{1}{\tau},~~~~ \langle a_{sf}( \tau ) a_{sf}(0) \rangle =\frac{\gamma^{2}_{2}}{\tau^{3}} \end{eqnarray} From the above equation, We can read the scaling dimensions of the various fields: $[\widehat{a}]=[\widehat{b}]=[a_{s}]=[b_{s}]=1/2, [a_{sf}]=[b_{sf}]=3/2 $. At the fixed point, the impurity degree of freedoms completely disappear: $\widehat{a}, \widehat{b} $ turn into the {\em non-interacting } scaling fields at the fixed point \begin{equation} \widehat{a} \sim b_{sf}(0,\tau),~~~ \widehat{b} \sim a_{sf}(0,\tau) \end{equation} Using Eqs.\ref{imp}, \ref{reverse}, the impurity spin turns into \begin{eqnarray} S_{x}(\tau) &= & i( -a_{sf}(0,\tau) b_{s}(0,\tau)+b_{sf}(0,\tau) a_{s}(0,\tau) ) \nonumber \\ S_{y}(\tau) &= & i( a_{sf}(0,\tau) a_{s}(0,\tau)+b_{sf}(0,\tau) b_{s}(0,\tau) ) \nonumber \\ S_{z}(\tau) &= & i ( a_{sf}(0,\tau) b_{sf}(0,\tau) + a_{s}(0,\tau) b_{s}(0,\tau) ) \end{eqnarray} The impurity spin-spin correlation function show typical FL behavior \begin{equation} \langle S^{z}(\tau) S^{z}(0) \rangle =\frac{1}{\tau^{2}} \end{equation} Using the fermionized form of the Eq.\ref{current} and paying attention to the {\em spinor} nature of the representation, it is easy to see the impurity spin renormalizs into either $ \vec{J}_{1} (0,\tau) $ or $ \vec{J}_{2}(0,\tau) $ depending on the sign of $\alpha $. This is consistent with the CFT analysis in the Appendix. There are 4 leading irrelevant operators with dimension 2 in the action $ S $ : $ \gamma_{1} $ and $ \gamma_{2} $ terms, $\lambda_{z}^{\prime} $ term and $ a_{s}(0,\tau) \frac{ \partial a_{s}(0,\tau)}{\partial \tau} + b_{s}(0,\tau) \frac{ \partial b_{s}(0,\tau)}{\partial \tau} $ which will be generated by the $\lambda_{z}^{\prime} $ term. The $ \alpha_{z} $ term has dimension 4, it can be written as $ :\widehat{a}(\tau) \frac{\partial \widehat{a}(\tau)}{\partial \tau}: :\widehat{b}(\tau) \frac{\partial \widehat{b}(\tau)}{\partial \tau}: $. The bosonized forms of the 4 leading irrelevant operators are \cite{another} \begin{eqnarray} \widehat{a}(\tau) \frac{\partial \widehat{a}(\tau)}{\partial \tau} & = & \cos 2\Phi_{sf}-\frac{1}{2} (\partial \Phi_{sf}(0))^{2} \nonumber \\ \widehat{b}(\tau) \frac{\partial \widehat{b}(\tau)}{\partial \tau} & = & -\cos 2\Phi_{sf}-\frac{1}{2} (\partial \Phi_{sf}(0))^{2} \nonumber \\ \widehat{a}\widehat{b}a_{s}(0)b_{s}(0) & = & \partial\Phi_{sf}(0,\tau) \partial\Phi_{s}(0,\tau) \nonumber \\ a_{s}(0,\tau) \frac{ \partial a_{s}(0,\tau)}{\partial \tau} & + & b_{s}(0,\tau) \frac{ \partial b_{s}(0,\tau)}{\partial \tau} = (\partial \Phi_{s}(0,\tau))^{2} \label{four} \end{eqnarray} Following the method developed in Ref.\cite{powerful}, their contributions to the single particle Green functions can be calculated. The first order correction to the single particle L-R Green function ( $ x_{1}>0, x_{2}<0 $ ) due to the first operator in the above Eq. is \begin{eqnarray} &\langle & \psi_{1 \uparrow}( x_{1},\tau_{1} ) \psi^{\dagger}_{1 \uparrow}( x_{2},\tau_{2} ) \rangle = \int d\tau \langle e^{-\frac{i}{2} \Phi_{c}( x_{1}, \tau_{1} )} e^{\frac{i}{2} \Phi_{c}( x_{2}, \tau_{2} )}\rangle \nonumber \\ & \times &\langle e^{-\frac{i}{2} \Phi_{s}( x_{1}, \tau_{1} )} e^{\frac{i}{2} ( \Phi_{s}( x_{2}, \tau_{2} ) + \pi )} \rangle \langle e^{-\frac{i}{2} \Phi_{f}( x_{1}, \tau_{1} )} e^{\frac{i}{2} \Phi_{f}( x_{2}, \tau_{2} )}\rangle \nonumber \\ & \times & \langle e^{-\frac{i}{2} \Phi_{sf}( x_{1}, \tau_{1} )} (:\cos2 \Phi_{sf}( 0, \tau ): -\frac{1}{2} : (\partial \Phi_{sf}( 0,\tau) )^{2} :) e^{\frac{i}{2} ( \Phi_{sf}( x_{2}, \tau_{2} ) +\pi )}\rangle \nonumber \\ & \sim & (z_{1}-\bar{z}_{2} )^{-2} \label{single} \end{eqnarray} Where $ z_{1}=\tau_{1}+i x_{1} $ is in the upper half plane, $ \bar{z}_{2} =\tau_{2}+i x_{2} $ is in the lower half plane. By using the following OPE: \begin{eqnarray} : e^{-\frac{i}{2} \Phi_{sf}( z_{1} )}: : e^{\frac{i}{2} \Phi_{sf}( z_{2})}: = (z_{1}-z_{2})^{-1/4}-\frac{i}{2}(z_{1}-z_{2})^{3/4} :\partial \Phi_{sf}(z_{2}): \nonumber \\ -\frac{i}{4}(z_{1}-z_{2})^{7/4} :\partial^{2} \Phi_{sf}(z_{2}): -\frac{1}{8}(z_{1}-z_{2})^{7/4} : (\partial \Phi_{sf}(z_{2}) )^{2}: + \cdots \label{ope} \end{eqnarray} It is ease to see that the primary field $ :\cos2 \Phi_{sf}( 0, \tau ): $ makes {\em no} contributions to the three point function. It was shown by the detailed calculations in Ref.\cite{conductivity} that only the part of the self-energy which is both {\em imaginary} and {\em even} function of $ \omega $ contributes to the conductivity. Although the energy momentum tensor $ : (\partial \Phi_{sf}( 0,\tau) )^{2} : $ do make $\sim \omega $ contribution to the self-energy in the first order \cite{conn}, because it is a {\em odd} function, it does {\em not} contribute to the electron conductivity in this order. Same arguments apply to the other operators in Eq.\ref{four}. Second order perturbations in these operators lead to the generic $ T^{2} $ fermi liquid bahaviour of the electron conductivity. The results of this section were applied to a two level tunneling system with slight modifications in Ref.\cite{hopping}. The universal scaling functions in the presence of external magnetic field which breaks the channel symmetry were also discussed there. \section{Compactified one channel Kondo Model} Assuming Particle-Hole symmetry, the {\em non-interacting} one channel Kondo model has two commuting $ SU(2) $ symmetry, one is the usual spin symmetry with the generators $ J^{a} (a=x,y,z) $ another is the isospin symmetry with the generators $ I^{a} (a=x,y,z) $. \begin{eqnarray} J_{x} & = & \frac{1}{2}( \psi^{\dagger}_{\uparrow} \psi_{\downarrow} + \psi^{\dagger}_{\downarrow} \psi_{\uparrow} ), ~~ J_{y}=\frac{1}{2i}( \psi^{\dagger}_{\uparrow} \psi_{\downarrow} - \psi^{\dagger}_{\downarrow} \psi_{\uparrow} ), ~~ J_{z}=\frac{1}{2}( \psi^{\dagger}_{\uparrow} \psi_{\uparrow} - \psi^{\dagger}_{\downarrow} \psi_{\downarrow} ) \nonumber \\ I_{x} & = & \frac{1}{2}( \psi^{\dagger}_{\uparrow} \psi^{\dagger}_{\downarrow} + \psi_{\downarrow} \psi_{\uparrow} ), ~~ I_{y}=\frac{1}{2i}( \psi^{\dagger}_{\uparrow} \psi^{\dagger}_{\downarrow} - \psi_{\downarrow} \psi_{\uparrow}), ~~ I_{z}=\frac{1}{2}( \psi^{\dagger}_{\uparrow} \psi_{\uparrow} + \psi^{\dagger}_{\downarrow} \psi_{\downarrow} ) \label{si} \end{eqnarray} The diagonal and off-diagonal components of the isospin currents represent respectively the charge and pairing density at the site $ x$. The one channel compactified model proposed by Ref.\cite{coleman} is a model where the impurity spin couples to both the spin and the isospin currents of the one channel conduction electrons \begin{eqnarray} H_{c} &= & i v_{F} \int^{\infty}_{-\infty} dx \psi^{\dagger}_{ \alpha }(x) \frac{d \psi_{ \alpha }(x)}{dx} + \sum_{a=x,y,z} \lambda^{a}( I^{a}(0)+ J^{a}(0) ) S^{a} + \sum_{a=x,y,z} \alpha^{a} (I^{a}(0)-J^{a}(0)) S^{a} \nonumber \\ & + & h ( \int dx (I^{z}(x) + J^{z}(x)) + S^{z} ) \label{com} \end{eqnarray} The ordinary symmetric Anderson impurity model in a one dimensional lattice is \begin{eqnarray} H & = & i t \sum_{n,\alpha} ( \psi^{\dagger}_{\alpha}(n+1) \psi_{\alpha}(n) - h. c. ) \nonumber \\ & + & i V \sum_{\alpha} ( \psi^{\dagger}_{\alpha}(0) d_{\alpha} -h.c.) + U(n_{d \uparrow}-\frac{1}{2})(n_{d \downarrow}-\frac{1}{2}) \label{aim} \end{eqnarray} The $ O(4) $ symmetry of the AIM can be clearly displayed in terms of the Majorana fermions \begin{eqnarray} \psi_{\uparrow}(n) & = &\frac{1}{\sqrt{2}} ( \chi_{1}(n) -i \chi_{2}(n) ), ~~~ d_{\uparrow}=\frac{1}{\sqrt{2}}( d_{1}-i d_{2} ) \nonumber \\ \psi_{\downarrow}(n) & = & \frac{1}{\sqrt{2}} ( -\chi_{3}(n) -i \chi_{0}(n) ), ~~~ d_{\downarrow}=\frac{1}{\sqrt{2}}( -d_{3}-i d_{0} ) \end{eqnarray} Breaking the symmetry from $ O(4) $ to $ O(3) \times O(1) $ in the hybridization \cite{ising}, the Hamiltonian \ref{aim} becomes: \begin{eqnarray} H & = & i t \sum_{n} \sum^{3}_{\alpha=0} \chi_{\alpha}(n+1) \chi_{\alpha}(n) +i V_{0} \chi_{0}(0) d_{0} \nonumber \\ & + & i V \sum^{3}_{ \alpha=1} \chi_{\alpha}(0) d_{\alpha} + U d_{1} d_{2} d_{3} d_{0} \label{break} \end{eqnarray} In the large $ U $ limit, projecting out the excited impurity states, we can map the Hamiltonian \ref{break} to the 1CCK Hamiltonian \ref{com} with \begin{equation} \lambda= \frac{ 2 V^{2}}{U},~~~ \alpha= -\frac{ 2 V_{0} V }{ U } \end{equation} If $ V_{0}=V $, Eq.\ref{break} comes back to the original $ O(4) $ symmetric AIM. In the strong coupling limit, it becomes the one channel Kondo model where the impurity only couples to the spin currents (or isospin currents) of the conduction electrons \cite{exchange}. If $ V_{0} =0 $, then $ \alpha=0 $, Eq.\ref{break} becomes the isotropic 1CCK where the impurity couples to the spin and isospin currents with equal strength. If we define the P-H transformation $ \psi_{\uparrow} \rightarrow \psi_{\uparrow}, \psi_{\downarrow} \rightarrow \psi^{\dagger}_{\downarrow} $, then spin and isospin currents transform to each other $ I^{a} \rightarrow J^{a}, J^{a} \rightarrow I^{a} $. The Hamiltonian \ref{com} has the P-H symmetry if $\alpha=0 $. In the following, parallel to the discussions on the 2CFAK, we take $ \lambda^{x}=\lambda^{y}=\lambda, \lambda^{z} \neq \lambda; \alpha^{x}=\alpha^{y}=\alpha, \alpha^{z} \neq \alpha $. We bosonize the spin $\uparrow $ and spin $\downarrow $ electrons separately \begin{equation} \psi_{ \alpha }(x )= \frac{P_{ \alpha}}{\sqrt{ 2 \pi a }} e ^{- i \Phi_{ \alpha}(x) } \label{one} \end{equation} The cocyle factors have been chosen as $ P_{ \uparrow}= P_{ \downarrow} = e^{i \pi N_{ \uparrow} } $. The bosonized form of the spin and isospin currents in Eq.\ref{si} are \begin{eqnarray} J_{x} & = & \frac{1}{2\pi a} \cos \sqrt{2} \Phi_{s},~~~ J_{y}=\frac{1}{2\pi a} \sin \sqrt{2} \Phi_{s},~~~ J_{z}=-\frac{1}{4\pi } \frac{\partial}{\partial x} \sqrt{2} \Phi_{s} \nonumber \\ I_{x} & = & \frac{1}{2\pi a} \cos \sqrt{2} \Phi_{c},~~~ I_{y}=\frac{1}{2\pi a} \sin \sqrt{2} \Phi_{c},~~~ I_{z}=-\frac{1}{4\pi } \frac{\partial}{\partial x} \sqrt{2} \Phi_{c} \end{eqnarray} where $ \Phi_{c}, \Phi_{s} $ are charge and spin bosons: \begin{eqnarray} \Phi_{c} = \frac{1}{\sqrt{2}}( \Phi_{\uparrow}+\Phi_{\downarrow}),~~~~~ \Phi_{s} = \frac{1}{\sqrt{2}}( \Phi_{\uparrow}-\Phi_{\downarrow}) \label{cs} \end{eqnarray} The sum $ J_{s}^{a}(x) = I^{a}(x) + J^{a}(x) $ and the difference $ J_{d}^{a}(x) = I^{a}(x) - J^{a}(x) $ can be expressed in terms of the chiral bosons \begin{eqnarray} J^{s}_{x}= \frac{1}{\pi a} \cos \Phi_{\uparrow} \cos \Phi_{\downarrow},~ J^{s}_{y}= \frac{1}{\pi a} \sin \Phi_{\uparrow} \cos \Phi_{\downarrow},~ J^{s}_{z}= -\frac{1}{ 2 \pi} \frac{ \partial \Phi_{\uparrow}}{\partial x} \nonumber \\ J^{d}_{x}=- \frac{1}{\pi a} \sin \Phi_{\uparrow} \sin \Phi_{\downarrow},~ J^{d}_{y}= \frac{1}{\pi a} \cos \Phi_{\uparrow} \sin \Phi_{\downarrow},~ J^{d}_{z}= -\frac{1}{ 2 \pi} \frac{ \partial \Phi_{\downarrow}}{\partial x} \label{iso} \end{eqnarray} Compare Eq.\ref{current} with Eq.\ref{iso}, we immediately realize that the mapping between the 2CFAK and the 1CCK is $ \Phi_{s} \rightarrow \Phi_{\uparrow}, \Phi_{sf} \rightarrow \Phi_{\downarrow} $, therefore $ \psi_{s} \rightarrow \psi_{\uparrow}, \psi_{sf} \rightarrow \psi_{\downarrow} $. The following fixed point structure can be immediately borrowed from the corresponding discussions on the 2CFAK. \subsection{ Fixed point 1 } The boundary conditions are \begin{equation} \psi_{\uparrow,L} = -\psi_{\uparrow,R},~~~ \psi_{\downarrow,L} =\psi^{\dagger}_{\downarrow,R} \end{equation} It is easy to see that the above boundary conditions respect the P-H symmetry, they can be expressed in terms of bosons \begin{equation} \Phi_{\uparrow,L} =\Phi_{\uparrow,R} + \pi,~~~ \Phi_{\downarrow,L} = -\Phi_{\downarrow,R} + \pi \end{equation} Spin $\uparrow $ electrons suffer a $ \frac{\pi}{2} $ phase shift, however, spin $\downarrow $ electrons are scattered into holes and vice-versa. The one particle S-matrix are $ S_{\uparrow}=-1, S_{\downarrow}=0 $. The residual conductivity of the spin $\uparrow$ electron takes unitary limit, but that of the spin $\downarrow $ is half of the unitary limit. The isotropic 1CCK has the same thermodynamic behaviors as the 2CK, but its fixed point has the local KM symmetry $ \widehat{O}_{1}(3) \times \widehat{O}_{1}(1) $. The finite size spectrum of this NFL fixed point is listed in Table \ref{compactnfl}. Comparing this finite size spectrum with that of the NFL fixed point of the 2CK listed in Ref.\cite{powerful}, it is easy to see that it has the {\em same } energy levels as those of the 2CK, but the corresponding degeneracy is {\em much smaller}. This is within the expectation, because the central charge $ c=2 $ and the fixed point symmetry of the isotropic 1CCK is smaller than that of the 2CK. This fixed point is stable only when $\alpha=\alpha_{z}=0 $ where the Hamiltonian \ref{com} has P-H symmetry. Away from the fixed point, there is only one dimension 3/2 operator \begin{equation} \widehat{a} \widehat{b} \partial \Phi_{\uparrow}(0) \sim \cos \Phi_{\downarrow}(0) \partial \Phi_{\uparrow}(0) \end{equation} The first order correction to the single particle L-R Green function ( $ x_{1}>0, x_{2}<0 $ ) due to this operator is \begin{eqnarray} \int d\tau \langle e^{- i \Phi_{\uparrow}( x_{1}, \tau_{1} )} \partial \Phi_{\uparrow}(0,\tau) e^{ i ( \Phi_{\uparrow}( x_{2}, \tau_{2} ) +\pi )}\rangle \langle :\cos \Phi_{\downarrow}( 0, \tau ): \rangle =0 \nonumber \\ \int d\tau \langle e^{- i \Phi_{\downarrow}( x_{1}, \tau_{1} )} \cos \Phi_{\downarrow}(0,\tau) e^{- i \Phi_{\downarrow}( x_{2}, \tau_{2} )}\rangle \langle \partial \Phi_{\uparrow}( 0, \tau ) \rangle =0 \end{eqnarray} By Wick theorem, it is easy to see that any {\em odd} order corrections vanish. Second order correction goes as $ \sim \omega $ which is a {\em odd} function, therefore does not contribute to the electron conductivity. The fourth order makes $ T^{2} $ contributions. There are two dimension 2 operators: \begin{eqnarray} a_{\uparrow}(0,\tau) \frac{ \partial a_{\uparrow}(0,\tau)}{\partial \tau} & + & b_{\uparrow}(0,\tau) \frac{ \partial b_{\uparrow}(0,\tau)}{\partial \tau} = (\partial \Phi_{\uparrow}(0,\tau))^{2} \nonumber \\ \widehat{a}(\tau) \frac{\partial \widehat{a}(\tau)}{\partial \tau} & = & \cos 2\Phi_{\downarrow}-\frac{1}{2} (\partial \Phi_{\downarrow}(0))^{2} \label{cool} \end{eqnarray} The first order correction to the spin $\uparrow $ electron L-R Green function due to the first operator in Eq.\ref{cool} is \begin{eqnarray} \int d\tau \langle e^{- i \Phi_{\uparrow}( x_{1}, \tau_{1} )} :( \partial \Phi_{\uparrow}(0,\tau) )^{2}: e^{ i ( \Phi_{\uparrow}( x_{2}, \tau_{2} ) +\pi )}\rangle \sim ( z_{1}-\bar{z}_{2} )^{-2} \label{many} \end{eqnarray} As pointed out in the last section, the energy momentum tensor $ : (\partial \Phi_{\uparrow}( 0,\tau) )^{2} : $ makes $\sim \omega $ contribution to the self-energy in the first order, therefore does not contribute to the electron conductivity. Second order perturbation in this operator leads to $ T^{2} $ contributions. Adding the contributions from all the leading irrelevant operators, we get \begin{equation} \sigma_{\uparrow}(T) \sim \sigma_{u}(1+ T^{2} + T^{4} + \cdots) \end{equation} The first order correction to the spin $\downarrow $ electron L-R Green function due to the 2nd operator in Eq. \ref{cool} is \begin{eqnarray} \int d\tau \langle e^{- i \Phi_{\downarrow}( x_{1}, \tau_{1} )} \cos 2\Phi_{\downarrow}(0,\tau) e^{- i \Phi_{\downarrow}( x_{2}, \tau_{2} )}\rangle \nonumber \\ -\frac{1}{2} \int d\tau \langle e^{- i \Phi_{\downarrow}( x_{1}, \tau_{1} )} ( \partial \Phi_{\downarrow}(0,\tau) )^{2} e^{- i \Phi_{\downarrow}( x_{2}, \tau_{2} )}\rangle \end{eqnarray} By using the following OPE: \begin{equation} : e^{-i \Phi_{\downarrow}( z_{1} )}: : e^{-i \Phi_{\downarrow}( z_{2})}: = (z_{1}-z_{2}): e^{-i 2 \Phi_{\downarrow}( z_{2} )}: -i(z_{1}-z_{2})^{2} : e^{-i 2 \Phi_{\downarrow}( z_{2} )} \partial \Phi_{\downarrow}(z_{2}): +\cdots \end{equation} It is ease to see that the {\em second} integral vanishes, but the {\em first} becomes \begin{equation} \frac{1}{ (z_{1}-\bar{z}_{2} )^{-1} } \int d \tau \frac{1}{ (z_{1}-\tau)^{2} (\tau-\bar{z}_{2} )^{2} } \sim (z_{1}-\bar{z}_{2} )^{-2} \end{equation} Putting $ \Delta=1 $ in Eq. (3.52) of Ref.\cite{conductivity}, we find the imaginary and real parts of self-energy go as $ Im \Sigma(\omega, T=0)=0, Re \Sigma(\omega, T=0) \sim \omega $, therefore the first order perturbation does not contribute to the spin $\downarrow $ electron conductivity. Second order perturbation yields a $ T^{2} $ contributions. Adding the contributions from all the leading irrelevant operators, we get \begin{equation} \sigma_{\downarrow}(T) \sim 2 \sigma_{u}(1+ T^{2}+ T^{4} + \cdots ) \end{equation} The total conductivity is the summation of the two spin components \cite{bhatt} \begin{equation} \sigma(T)= \sigma_{\uparrow}(T)+ \sigma_{\downarrow}(T) \sim 3 \sigma_{u}(1+ T^{2}+ T^{4} + \cdots ) \end{equation} The boundary OPE of the spin and density of the {\em conduction electrons} are \begin{eqnarray} \psi^{\dagger}_{\uparrow}(z_{1}) \psi_{\uparrow}( \bar{z}_{2} ) & = & ( z_{1}-\bar{z}_{2} )^{-1} +i \partial \Phi_{\uparrow} + \cdots \nonumber \\ \psi^{\dagger}_{\downarrow}(z_{1}) \psi_{\downarrow}( \bar{z}_{2} ) & = & 0 + \cdots \nonumber \\ \psi^{\dagger}_{\uparrow}(z_{1}) \psi_{\downarrow}( \bar{z}_{2} ) & = & e^{i \sqrt{2} \Phi_{c}(0) } + \cdots \nonumber \\ \psi^{\dagger}_{\downarrow}(z_{1}) \psi_{\uparrow}( \bar{z}_{2} ) & = & e^{-i \sqrt{2} \Phi_{s}(0) } + \cdots \label{sdnfl} \end{eqnarray} The boundary OPE of the spin singlet and triplet pairing operators are \begin{eqnarray} \psi_{\uparrow}(z_{1}) \psi_{\uparrow}( \bar{z}_{2} ) & = & 0 + \cdots \nonumber \\ \psi_{\downarrow}(z_{1}) \psi_{\downarrow}( \bar{z}_{2} ) & = & ( z_{1}-\bar{z}_{2} )^{-1} -i \partial \Phi_{\downarrow} + \cdots \nonumber \\ \psi_{\uparrow}(z_{1}) \psi_{\downarrow}( \bar{z}_{2} ) & = & e^{-i \sqrt{2} \Phi_{s}(0) } + \cdots \nonumber \\ \psi_{\downarrow}(z_{1}) \psi_{\uparrow}( \bar{z}_{2} ) & = & - e^{-i \sqrt{2} \Phi_{c}(0) } + \cdots \label{pairingnfl} \end{eqnarray} The P-H symmetry interchanges the pairing and spin operators in the $ \uparrow \downarrow $ and $ \downarrow \uparrow $ channels. From Eq.\ref{pairingnfl}, we can identify the pairing operators \begin{equation} {\cal O}_{s}= e^{-i \sqrt{2} \Phi_{s}(0) },~~~ {\cal O}_{c}= e^{-i \sqrt{2} \Phi_{c}(0)},~~~ {\cal O}_{\downarrow}= \partial \Phi_{\downarrow}(0) \end{equation} The paring operators in all the channels except in the $ \uparrow \uparrow $ channel have scaling dimension 1, therefore their correlation functions decay as $ \tau^{-2} $. Comparing these pairing operators with those at the FL fixed point ( Eq. \ref{pairingfl} ) to be discussed in the following, we find the pairng susceptibility in $ \downarrow \downarrow $ channel is enhanced. However, in contrast to the 2CK fixed point \cite{powerful}, the enhancement is so weak that there is {\em no} pairing susceptibility {\em divergence} at the impurity site in {\em any spin channel}. This result is somewhat surprising. Naively, we expect pairing susceptibility divergence because the impurity interacts with the pairing density of the conduction electrons at the impurity site. However, the above explicit calculations showed that this is {\em not} true if there is only {\em one } channel of conduction electrons. Naively, we do {\em not} expect pairing susceptibility divergence in the 2CK, because the impurity spin interacts only with the total {\em spin } currents of channel 1 and 2, {\em no } isospin currents of channel 1 and 2 are involved in the interaction. However, the explicit calculation of the 2CK showed that the pairing operator in the spin and flavor singlet channel has dimension 1/2 ( however, the pairing operators in flavor singlet and spin triplet channel has dimension 3/2 ), therefore the spin and flavor singlet pairing susceptiblity at the impurity site is {\em divergent} \cite{powerful}. This indicates that we can achieve the pairing susceptiblity divergence without a pairing source term. We conclude that {\em more than } one channel of conduction electrons are needed to achieve the pairing susceptipility {\em divergence}. \subsection{ Fixed pointed 2} The boundary conditions are \begin{equation} \psi_{\uparrow,L} = -\psi_{\uparrow,R},~~~ \psi_{\downarrow,L} = -\psi^{\dagger}_{\downarrow,R} \end{equation} The above boundary conditions can be expressed in terms of bosons \begin{equation} \Phi_{\uparrow,L} =\Phi_{\uparrow,R} + \pi,~~~ \Phi_{\downarrow,L} = -\Phi_{\downarrow,R} \end{equation} This fixed point is stable only when $\lambda=\alpha_{z}=0 $. If we define the P-H transformation $ \psi_{\uparrow} \rightarrow \psi_{\uparrow}, \psi_{\downarrow} \rightarrow -\psi^{\dagger}_{\downarrow} $, then the spin and isospin currents transform as $ I^{x} \rightarrow -J^{x}, I^{y} \rightarrow -J^{y}, I^{z} \rightarrow J^{z}; J^{x} \rightarrow -I^{x}, J^{y} \rightarrow -I^{y}, J^{z} \rightarrow I^{z}$. The Hamiltonian \ref{com} has this P-H symmetry if $\lambda=\alpha_{z}=0 $. This is the same fixed point as fixed point 1. \subsection{ Fixed pointed 3} The boundary conditions are \begin{equation} \psi_{\uparrow,L} =-\psi_{\uparrow,R},~~~ \psi_{\downarrow,L} = -\psi_{\downarrow,R} \end{equation} The above boundary conditions can be expressed in terms of bosons \begin{equation} \Phi_{\uparrow,L} =\Phi_{\uparrow,R} + \pi,~~~ \Phi_{\downarrow,L} = \Phi_{\downarrow,R} + \pi \end{equation} Both spin $\uparrow $ and $\downarrow $ electrons suffer $ \frac{\pi}{2} $ phase shift. The physical picture is that the impurity spin is either totally screened by the spin current or the isospin current of conduction electrons depending on which coupling is stronger \cite{exchange}. This is a FL fixed point with $ O(4) $ symmetry. The finite size spectrum is listed in Table \ref{compactfl}. The bosonized forms of the 4 leading irrelevant operators are \cite{another} \begin{eqnarray} \widehat{a}(\tau) \frac{\partial \widehat{a}(\tau)}{\partial \tau} & = & \cos 2\Phi_{\downarrow}-\frac{1}{2} (\partial \Phi_{\downarrow}(0))^{2} \nonumber \\ \widehat{b}(\tau) \frac{\partial \widehat{b}(\tau)}{\partial \tau} & = & -\cos 2\Phi_{\downarrow}-\frac{1}{2} (\partial \Phi_{\downarrow}(0))^{2} \nonumber \\ \widehat{a}\widehat{b}a_{\uparrow}(0)b_{\uparrow}(0) & = & \partial\Phi_{\downarrow}(0,\tau) \partial\Phi_{\uparrow}(0,\tau) \nonumber \\ a_{\uparrow}(0,\tau) \frac{ \partial a_{\uparrow}(0,\tau)}{\partial \tau} & + & b_{\uparrow}(0,\tau) \frac{ \partial b_{\uparrow}(0,\tau)}{\partial \tau} = (\partial \Phi_{\uparrow}(0,\tau))^{2} \label{last} \end{eqnarray} The first order correction to the spin $\uparrow $ electron L-R Green function due to the 4th operator in Eq.\ref{last} is also given by Eq.\ref{many}. The correction due to the 3rd operator in Eq.\ref{last} can be similarly evaluated. We get the low temperature expansion of the spin $\uparrow $ electron conductivity \begin{equation} \sigma_{\uparrow}(T) \sim \sigma_{u}(1+ T^{2} + T^{4} + \cdots) \label{phyup} \end{equation} The first order correction to the spin $\downarrow $ electron L-R Green function due to the first operator in Eq. \ref{last} is \begin{eqnarray} \int d\tau \langle e^{- i \Phi_{\downarrow}( x_{1}, \tau_{1} )} \cos 2\Phi_{\downarrow}(0,\tau) e^{ i ( \Phi_{\downarrow}( x_{2}, \tau_{2} ) +\pi) }\rangle \nonumber \\ -\frac{1}{2} \int d\tau \langle e^{- i \Phi_{\downarrow}( x_{1}, \tau_{1} )} ( \partial \Phi_{\downarrow}(0,\tau) )^{2} e^{ i ( \Phi_{\downarrow}( x_{2}, \tau_{2} ) +\pi)}\rangle \end{eqnarray} By using the following OPE: \begin{eqnarray} : e^{-i \Phi_{\downarrow}( z_{1} )}: : e^{i \Phi_{\downarrow}( z_{2})}: = (z_{1}-z_{2})^{-1}-i :\partial \Phi_{\downarrow}(z_{2}): \nonumber \\ +\frac{z_{1}-z_{2}}{2} :\partial^{2} \Phi_{\downarrow}(z_{2}): -\frac{z_{1}-z_{2}}{2} : (\partial \Phi_{\downarrow}(z_{2}) )^{2}: + \cdots \end{eqnarray} It is ease to see that the first integral vanishes and the second are the same as Eq.\ref{many}. The corrections due to the 2nd and the 3rd operators in Eq.\ref{last} can be similarly evaluated, the low temperature expansion of the spin $\downarrow $ electron conductivity follows \begin{equation} \sigma_{\downarrow}(T) \sim \sigma_{u}(1+ T^{2} + T^{4} + \cdots) \label{phydown} \end{equation} Note that only at the FL fixed point, the spin $ SU(2) $ symmetry is restored, therefore the expansion coefficients in Eqs.\ref{phyup},\ref{phydown} are {\em different}. The total conductivity is the summation of the two spin components \begin{equation} \sigma(T)= \sigma_{\uparrow}(T)+ \sigma_{\downarrow}(T) \sim 2 \sigma_{u}(1+ T^{2}+ T^{4} + \cdots ) \end{equation} The boundary OPE of the spin and density of the {\em conduction electrons} are \begin{eqnarray} \psi^{\dagger}_{\uparrow}(z_{1}) \psi_{\uparrow}( \bar{z}_{2} ) & = & ( z_{1}-\bar{z}_{2} )^{-1} +i \partial \Phi_{\uparrow} + \cdots \nonumber \\ \psi^{\dagger}_{\downarrow}(z_{1}) \psi_{\downarrow}( \bar{z}_{2} ) & = & ( z_{1}-\bar{z}_{2} )^{-1} +i \partial \Phi_{\downarrow} + \cdots \nonumber \\ \psi^{\dagger}_{\uparrow}(z_{1}) \psi_{\downarrow}( \bar{z}_{2} ) & = & e^{i \sqrt{2} \Phi_{s}(0) } + \cdots \nonumber \\ \psi^{\dagger}_{\downarrow}(z_{1}) \psi_{\uparrow}( \bar{z}_{2} ) & = & e^{-i \sqrt{2} \Phi_{s}(0) } + \cdots \label{sdfl} \end{eqnarray} The boundary OPE of the spin singlet and triplet pairing operators are \begin{eqnarray} \psi_{\uparrow}(z_{1}) \psi_{\uparrow}( \bar{z}_{2} ) & = & 0 + \cdots \nonumber \\ \psi_{\downarrow}(z_{1}) \psi_{\downarrow}( \bar{z}_{2} ) & = & 0 +\cdots \nonumber \\ \psi_{\uparrow}(z_{1}) \psi_{\downarrow}( \bar{z}_{2} ) & = & -e^{-i \sqrt{2} \Phi_{c}(0) } + \cdots \nonumber \\ \psi_{\downarrow}(z_{1}) \psi_{\uparrow}( \bar{z}_{2} ) & = & e^{-i \sqrt{2} \Phi_{c}(0) } + \cdots \label{pairingfl} \end{eqnarray} The above equations should be compared with the corresponding Eqs.\ref{sdnfl} and \ref{pairingnfl} at the NFL fixed point. \section{Conclusions} By the detailed discussions on the low temperature properties of the two related, but different single impurity models, we clarify the confusing conjectures and claims made on these two models. In evaluating the single particle Green functions and pairing susceptibilities, all the degree of freedoms have to be taken into account, even though some of them decouple from the interactions with the impurity. We explicitly demonstrate that different quantum impurity models are simply free chiral bosons with different boundary conditions. In Ref.\cite{sf}, the author studied another single impurity model where the impurity couples to both the spin and the flavor currents of the two channel electrons ( 2CSFK). In Ref.\cite{hopping}, the author solved a two level tunneling model which can also mapped to a single impurity model. As shown in Ref.\cite{twoimp}, finite number of impurity models can always mapped to a single impurity model. From the results of this paper and Refs.\cite{sf,hopping}, we conclude that in clean, finite number of impurity models (1) FL behaviors are extremely robust, any perturbation in the flavor sectors will destroy the NFL behaviors.(2) due to the phase space arguments given in this paper and in Refs.\cite{sf,hopping}, it is very unlikely to find the NFL linear $ T $ bahaviour of the electron conductivity which was observed in the certain heavy fermion systems \cite{linear} and in the normal state of high- $T_{c} $ cuprate superconductors. There are three possible ways to explain this experimental observation (1) disorder effects \cite{bhatt} (1) Kondo lattice model \cite{lattice} (3) near to some quantum phase transitions \cite{millis}, for example, near the phase transition between the metallic spin-glass and disordered metal \cite{phase,sachdev}. \centerline{\bf ACKNOWLEDGMENTS} We thank D. S. Fisher, B. Halperin, A. Millis, N. Read for helpful discussions. This research was supported by NSF Grants Nos. DMR 9630064, DMR9416910 and Johns Hopkins University.
proofpile-arXiv_065-415
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction and results} In this paper we investigate the problem of finding the mapping of a three-dimensional free fermion theory with non-abelian symmetry onto an equivalent bosonic quantum field theory. This mapping, commonly called bosonization, has been already established along the lines of the present investigation for the case of abelian symmetry \cite{FS}-\cite{LNS} and it has been also discussed in the non-abelian case, using a related method, in \cite{B}. In all these investigations we employ an approach close to that put forward in a series of very interesting works on smooth bosonization and duality bosonization \cite{DNS1}-\cite{DNS5}. Other related or alternative approaches to bosonization in $d>2$ dimensions have been also developed \cite{TSSG}-\cite{BBG}. The advantage of the bosonization method that we employ here lies on the fact that it provides a systematic procedure for deriving $d \ge 2$ bosonization recipes both for abelian and non-abelian symmetries. In this way, it gives an adequate framework for obtaining the bosonic equivalent of the original fermionic theory, the recipe for mapping fermionic and bosonic currents as well as the current commutation relations which are at the basis of bosonization. The approach we follow starts from the path-integral defining the generating functional for a theory of free fermions (including sources for fermionic currents) and ends with the generating functional for an equivalent bosonic theory. This allows to identify, {\it exactly}, the bosonization recipe for fermion currents independently of the number of space-time dimensions. Interestingly enough, one follows a series of steps which are the same for any space-time dimension and both for abelian and non-abelian symmetries. Of course, apart from the two-dimensional case and from the large fermion mass limit in the three dimensional case, one only achieves in general a partial bosonization in the sense that one cannot compute exactly the fermion path-integral in order to derive a local bosonic Lagrangian. Extending the three dimensional abelian bosonization approach discused in \cite{FS}-\cite{LNS} we derive in the present paper three dimensional non-abelian bo\-so\-niza\-tion. Concerning the fermion current, our method allows to derive the exact bosonization recipe \begin{equation} \bar \psi^i \gamma_\mu t^a_{ij} \psi^j \to \pm \frac{i}{8\pi}\varepsilon_{\mu \nu \alpha} \partial_\nu A^a_\alpha \,\, , \label{Ssss} \end{equation} where $i,j=1,\cdots,N$, $a=1,\cdots,{\rm dim}G$, $t^a$ are the generators and $f^{a b c}$ the structure constants of the symmetry group G. Finally $A_\mu$ is a vector field taking values in the Lie algebra of $G$. The knowledge of the bosonic action accompanying this bosonization rule is necessarily approximate since it implies the evaluation of the $d=3$ determinant for fermions coupled to a vector field. We then consider the case of very massive free fermions showing that, within this approximation, the fermion lagrangian bosonizes to the non-abelian Chern-Simons term, \begin{equation} \bar\psi^i \left(i\gamma^{\mu} \partial_{\mu}+m \right) \psi^i \to \mp \frac{i}{8 \pi} \epsilon_{\mu \nu\alpha} \left( A_{\mu}^a \partial_{\nu}A^a_{\alpha} + 1/3 f^{a b c} A_{\mu}^a A_{\nu}^b A_{\alpha}^c \right) ~, \label{aq11} \end{equation} This result, advanced in \cite{B} using a completely different approach, is the natural extension of the abelian result \cite{FS}-\cite{LNS}. In this last case the bosonic theory corresponds, in the large fermion mass limit, to a Chern-Simons theory while in the massless case it coincides with the abelian non-local action discussed in \cite{Mar},\cite{BFO}. In fact one should expect that an analysis similar to that in \cite{BFO} can be carried out leading to an explicit (although complicated) bosonic action valid in all fermion mass regimes. The results above are derived in section 3. As a warming up exercise, we rederive in section 2 the bosonization rules for two-dimensional non-abelian models. Indeed, following an approach related to that developed in \cite{BQn} for two-dimensional bosonization we arrive to the Wess-Zumino-Witten action and the well-known bosonization recipe for fermion currents \begin{equation} j_+ \to \frac{i}{4\pi}a^{-1} \partial_+ a \label{ele1} \end{equation} \begin{equation} j_- \to \frac{i}{4\pi}a \partial_- a^{-1} \label{ele2} \end{equation} with $a$ the group-valued bosonic field with dynamics governed by the Wess-Zumino-Witten action. Now, these rules, as well as the Polyakov-Wiegmann identity that we repeatedly use in its derivation, deeply relay on the holomorphic properties of two-dimensional theories \cite{PWI}\cite{Wi0}. Strikingly, we found that in the three dimensional case, a BRST symmetry structure underlying the bosonic version of the fermionic generating functional plays a similar role and allows to end, at least in the large mass limit, with a simple bosonization rule. This BRST symmetry is highly related to that used in \cite{DNS1}-\cite{DNS3}, \cite{DNS4}-\cite{TSSG}, and is analogous to that arising in topological field theories \cite{BS}-\cite{BRT}, its origin being related to the way the originally ``trivial'' bosonic field enters into play. \section{Warming up: $d=2$ non-abelian bosonization} Non-abelian bosonization in two dimensional space time was formulated first by Witten \cite{Wi} by comparing the current algebra for free fermions and for a bosonic sigma model with a Wess-Zumino term. Afterwards, different approaches rederived and discussed the bosonization recipe \cite{dVR}-\cite{GR} but in general they were not constructive in the sense that the bosonic theory was not obtained from the fermionic one by following a series of steps that could be generalized to other cases, in particular to a possible higher-dimensional bosonization. More recently, using the the duality technique \cite{BQ}-\cite{BLQ} which has as starting point the smooth bosonization approach \cite{DNS1}-\cite{DNS3}, \cite{DNS4}-\cite{DNS5} the recipe for non-abelian bosonization was obtained by Burgess and Quevedo \cite{BQn} in a way which is more adaptable to generalizations to higher dimensions. The approach to two-dimensional bosonization that we present in this section is related to that in \cite{BQn} and we think that it is worthwhile to describe it here in detail since it provides many of the clues which allow us to derive the non-abelian bosonization recipe for 3-dimensional fermions. Following Witten \cite{Wi} one can see that the bosonic picture for a theory of $N$ free massless Dirac fermions corresponds to a bosonic field $a \in SU(N)$ with a Wess-Zumino-Witten action and a real bosonic field $\phi$ with a free scalar field action \cite{dVR}-\cite{GO}. Since we shall be mainly interested in the specific non-abelian aspect of bosonization, we will not discuss the $\phi$ sector of the corresponding bosonic theory (although the method can trivially take into account the $U(1)$ sector associated with it). We start from the (Euclidean) Lagrangian for free massless Dirac fermions in $2$ dimensions \begin{equation} L = \bar\psi (\id) \psi \label{L} \end{equation} where fermions are in the fundamental representation of some group $G$. The corresponding generating functional reads \begin{equation} Z_{fer}[s] = \int D\bar \psi D\psi \exp[-\int d^2x \bar\psi (\id + /\kern-.52em s) \psi] \label{Zs} \end{equation} with $s_\mu = s_\mu^a t^a$ an external source taking values in the Lie algebra of $SU(N)$. Our derivation of bosonization rules both in $2$ and $3$ dimensions heavily relies on the invariance of the measure under local transformations of the fermion variables, $ \psi \to h(x) \psi$, $\bar \psi \to \bar \psi h(x)^{-1}$ with $h \in G$. As a consequence the generating functional (\ref{Zs}) is automatically invariant under local transformations of the source \begin{equation} s_\mu \to s_\mu^h = h^{-1} s_\mu h + i h^{-1} \partial_\mu h \label{inva} \end{equation} \begin{equation} Z[s^h] = Z[s] \label{idiss} \end{equation} In view of this, if we perform the change of variables \[ \psi =g(x) \psi' \] \begin{equation} \bar \psi = \bar \psi' g^{-1}(x) . \label{chas22} \end{equation} $Z_{fer}[s]$ becomes \begin{equation} Z_{fer}[s] = \int {\cal{D}} \bar \psi {\cal{D}} \psi {\cal{D}} g \exp[-\int d^2x \bar \psi (\id + /\kern-.52em s ^g ) \psi] \label{n2s} \end{equation} where an integration over $g$ has been included since it just amounts to a change of normalization. Integrating out fermions we have \begin{equation} Z_{fer}[s] = \int {\cal{D}} g \;\det(\id + {/\kern-.52em s}^g) \label{cambios} \end{equation} Now, posing \begin{equation} {s}_\mu^g = b_\mu \label{nuezz} \end{equation} and using \begin{equation} f_{\mu \nu}[b] = g^{-1} f_{\mu \nu}[s] g \label{rela} \end{equation} we shall trade the $g$ integration for an integration over connections $b$ satisfying condition (\ref{rela}). To this end we shall use the $d=2$ identity (proven in the Appendix) \begin{equation} \int {\cal{D}} b_\mu {\cal H}[b] \delta[\varepsilon_{\mu \nu} (f_{\mu\nu}[b] - f_{\mu\nu}[s])] = \int {\cal{D}} g {\cal H}[s^g] \label{idle} \end{equation} Here ${\cal H}$ is a gauge invariant function. Identity (\ref{idle}) allows us to write eq.(\ref{cambios}) in the form \begin{equation} Z_{fer} = \int {\cal{D}} b_\mu \Delta \delta(b_+ - s_+) \delta\left[\varepsilon_{\mu\nu}(f_{\mu \nu}[b]- f_{\mu \nu}[s])\right] \det(\id + /\kern-.52em b) \, . \label{n3ss} \end{equation} For convenience, we have chosen to fix the gauge using the condition $b_+ = s_+$ being $\Delta$ the corresponding Faddeev-Popov determinant. We now introduce a Lagrange multiplier ${\hat a}$ (taking values in the Lie algebra of $G$) to enforce the delta function condition \begin{eqnarray} Z_{fer}[s] & = & \int {\cal{D}} {\hat a}{\cal{D}} b_\mu \,\Delta \delta(b_+ - s_+) \det[\id +/\kern-.52em b] \times \nonumber \\ & & \exp \left(-\frac{C}{8\pi} tr \int d^2x {\hat a} \,\varepsilon _{\mu \nu } (f_{\mu \nu }[b] - f_{\mu \nu }[s])\right) \label{copia} \end{eqnarray} with $C$ a constant to be conveniently adjusted. We now write sources and the $b_\mu$ field in terms of group-valued variables, \begin{equation} s_+ = i {\tilde s}^{-1} \partial_+ \tilde s \label{s+} \end{equation} \begin{equation} s_- = i s \partial_- s^{-1} \label{s-} \end{equation} \begin{equation} b_+ = i (\tilde b \tilde s)^{-1} \partial_+ (\tilde b \tilde s) \label{b+} \end{equation} \begin{equation} b_- = i (sb) \partial_- (sb)^{-1} \label{ga} \end{equation} so that the fermion determinant can be related to the Wess-Zumino-Witten action \cite{PWI}, \begin{equation} \det[\id +/\kern-.52em b] = \exp (W[\tilde b \tilde s s b]) \label{deter} \end{equation} In writing eq.(\ref{deter}) a gauge-invariant regularization is assumed so that the left and right-handed sectors enter in gauge invariant combinations. In this way, gauge transformations of the source $s_\mu$, which as stated before, should leave the generating functional invariant, do not change the determinant. This will be the criterion we shall adopt each time determinants (always needing a regularization) has to be computed. Concerning the Jacobian for passing from the $b_\mu$ variable to the $b, \tilde b$ one can easily show that \begin{equation} \int \Delta \delta(b_+ - s_+){\cal{D}} b_\mu = \int \exp(\kappa W[\tilde b \tilde s s b]) \delta( \tilde b - I) {\cal{D}} \tilde b {\cal{D}} b \label{jaci} \end{equation} so that $Z_{fer}[s]$ becomes \begin{eqnarray} Z_{fer}[s] & = & \int {\cal{D}} {\hat a} {\cal{D}} b \exp \left(i\frac{C}{4\pi} tr \int d^2x ( D_+[\tilde s] {\hat a}) \, s b ( \partial_- b^{-1}) s^{-1}\right) \times \nonumber\\ & & \exp \left((1+\kappa)W[ \tilde s s b] \right) \, . \label{copiag} \end{eqnarray} A convenient change of variables to pass from integration over the algebra valued Lagrange multiplier ${\hat a}$ to a group valued variable $a$ is the one defined through \begin{equation} D_+[\tilde s] {\hat a} = {\tilde s}^{-1} (a^{-1}\partial_+a) \tilde s \label{nuez} \end{equation} Calculation of the corresponding jacobian $J_{L} $ \begin{equation} {\cal{D}} \hat a = J_{L} {\cal{D}} a \end{equation} should be carefully done. Indeed, in the present approach to bosonization, it is the group valued variable $a$ who plays the role of the boson field equivalent to the original fermion field. Since for the latter (a free fermi field) there was no local symmetry, the former should not be endowed with this symmetry. With this in mind, we shall maintain $a$ unchanged under local transformations $g$ while transforming $\hat a$, $\hat a \to g^{-1} \hat a g$, so that eq.(\ref{nuez}) changes covariantly when one simultaneously changes $\tilde s \to \tilde s g$. One can easily prove that \begin{equation} J_L = \exp (\kappa W[a \tilde s s] - \kappa W[\tilde s s]) \label{nuezi} \end{equation} so that the generating functional reads \begin{eqnarray} Z_{fer}[s] & = \int {\cal{D}} a {\cal{D}} b \exp((1 +\kappa)W[\tilde s s b]) \times \exp \left(\kappa (W[a \tilde s s]- W[\tilde s s])\right) \times \nonumber\\ &\exp(-\frac{C}{4\pi} tr \int d^2x {\tilde s}^{-1} (a^{-1} \partial_+ a) \tilde s s (b \partial_- b^{-1}) s^{-1} ) \, . \label{copg} \end{eqnarray} If one repeatedly uses Polyakov-Wiegmann identity and chooses the up to now arbitrary constant $C$ as \begin{equation} -C = 1 + \kappa \label{D} \end{equation} one can write $Z_{fer}[s]$ in the form \begin{equation} Z_{fer}[s] = \int {\cal{D}} a {\cal{D}} b \; \exp (W[\tilde s s] - W[ a \tilde s s]) \exp\left((1+\kappa)W{[a \tilde s s b ]} \right) \label{fifi} \end{equation} Now, the $b$ integration can be trivially factorized this leading to \begin{equation} Z_{fer}[s] = \int {\cal{D}} a \;\exp(-W[a \tilde s s] + W[\tilde s s] ) \,. \label{su} \end{equation} or, after the shift $ a {\tilde s} s \to \tilde s a s$ \label{ayy} \begin{eqnarray} Z_{fer}[s_+]& = & \int {\cal{D}} a \;\exp(-W[a] + \frac{i}{4\pi} tr \int d^2x (s_+ a \partial_- a^{-1} + \nonumber \\ & & s_- a^{-1} \partial_+ a) ) \times \exp(\frac{1}{4\pi} tr \int d^2x (a^{-1} s_+ a s_- - s_+ s_-)) \label{chuchi} \end{eqnarray} We have then arrived to the identity \begin{equation} Z_{fer}[s] = Z_{bos}[s] \label{cg} \end{equation} where $Z_{bos}[s]$ is the generating function for a Wess-Zumino-Witten model. Differentiation with respect to any one of the two sources gives correlation functions in a given chirality sector. The answer corresponds to Witten's bosonization recipe \cite{Wi} \begin{equation} \bar \psi t^a \gamma_+ \psi \to \frac{i}{4\pi} a^{-1} \partial_+ a \label{rec1} \end{equation} \begin{equation} \bar \psi t^a \gamma_- \psi \to \frac{i}{4\pi} a \partial_- a^{-1} \, , \label{rec2} \end{equation} the l.h.s. to be computed in a free fermionic model, the r.h.s. in a Wess-Zumino-Witten model. \section{ $d=3$ non-abelian bosonization} Contrary to the case of two-dimensional massless fermions, one cannot compute exactly the Dirac operator determinant for $d > 2$ in the presence of an arbitrary gauge field $b_\mu$, neither in the massless nor in the massive case. This implies the necessity of making approximations at some stage of our bosonization procedure to render calculations feasible. In the $d=3$ Abelian case one can handle these approximations in a very general framework \cite{BFO},\cite{LNS}. Being the non-Abelian case far more complicated than the Abelian one, we shall only discuss the limiting case of very massive fermions, for which the fermion determinant is related to the Chern-Simons (CS) action \cite{NS}-\cite{Red}. A second problem arising when one tries to extend the non-Abelian boson-fermion mapping from $d=2$ to $d=3$ concerns the central role that plays the Polyakov-Wiegman identity, related to the holomorphic properties of the two-dimensional model \cite{Wit}. In principle, a $3$-dimensional analogue of this identity is not available and this forbids a trivial extension to $d=3$ of the procedure described in the precedent section for two-dimensional bosonization. However, as we shall see, once one introduces the auxiliary field $b_\mu$, a BRST invariance of the kind arising in topological field theories \cite{BS}-\cite{BRT} can be unraveled. The use of BRST technique for bosonization of fermion models was initiated in the developement of the smooth bosonization approach \cite{DNS1}-\cite{DNS3}, \cite{DNS4}-\cite{DNS5}, closely related to bosonization duality and to the present treatment. In the present case, it allows to factor out the auxiliary field in the same way Polyakov-Wiegmann identity did the job in $d=2$. The resulting bosonization action coincides with that obtained using a completely different approach \cite{B}, based in the use of an interpolating Lagrangian \cite{VN}-\cite{vN}. The advantage of the present method lies in the fact that the BRST symmetry can be formulated in arbitrary dimensions while the interpolating Lagrangian, which replaces the role of this symmetry in decoupling auxiliary and bosonic fields is in principle applicable only in odd-dimensional spaces. We consider $N$ massive Dirac fermions in $d=3$ Euclidean dimensions with Lagrangian \begin{equation} L = \bar\psi (\id + m ) \psi \label{3L} \end{equation} The corresponding generating functional reads \begin{equation} Z_{fer}[s] = \int {\cal{D}} \bar \psi {\cal{D}} \psi \exp[-\int d^3x \bar \psi (\id + /\kern-.52em s + m ) \psi ] \label{41} \end{equation} Again, we introduce an auxiliary vector field $b_\mu$ and use the $d=3$ identity (proven in the Appendix) \begin{equation} Z_{fer}[s] = X[s]^{-1}\!\! \int \!\!Db_\mu X[b] \det(2\varepsilon_{\mu \nu \alpha} D_\nu[b]) \delta( {^*\!\!f}_\mu[b] - {^*\!\!f}_\mu[s] ) \det(\id + m + /\kern-.52em b) \label{IX} \end{equation} Here \begin{equation} ^*\!\!f_\mu = \varepsilon_{\mu \nu \alpha} f_{\nu\alpha} \label{II} \end{equation} Concerning $X[b]$, it is an arbitrary functional which can be introduced in order to control the issue of symmetries at each stage of our derivation. Indeed, bosonization of three dimensional very massive fermions ends with a bosonic field with dynamics governed by a Chern-Simons action. As explained in \cite{FAS}, an appropriate choice of $X$ allows to end with the natural gauge connection transformation law for this bosonic field. Following \cite{FAS}, we choose $X$ in the form \begin{equation} X[b] = \exp(\mp \frac{i}{24\pi} \varepsilon_{\mu\nu\alpha} tr \int d^3x b_\mu b_\nu b_\alpha ) \label{x} \end{equation} We can see at this point how an exact bosonization rule for the fermion current can be derived independently of the fact that one cannot calculate exactly the fermion determinant for $d > 2$. Indeed, if we introduce a Lagrange multiplier $A_\mu$ to represent the delta function, we can write $Z_{fer}$ in the form \begin{equation} Z_{fer}[s] = X[s]^{-1} \int {\cal{D}} A_\mu \exp \left( \mp \frac{i}{16\pi} tr \int d^3x A_\mu {^*\!\!f}_\mu[s] \right) \times \exp(-S_{bos}[A]) \label{ches} \end{equation} where we have defined the bosonic action $S_{bos}[A]$ as \begin{eqnarray} \exp(-S_{bos}[A]) & = & \int {\cal{D}} b_\mu \det(\id + m + /\kern-.52em b) X[b] \times \nonumber \\ & & \det(2\varepsilon_{\mu \nu \alpha}D_\nu[b]) \exp\left( \pm \frac{i}{16\pi} tr \int d^3x A_\mu {^*\!\!f}_\mu[b] \right) \label{jeje} \end{eqnarray} With the choice (\ref{x}) one indeed has gauge invariance of $S_{bos}[A]$, $A_\mu$ and $b_\mu$ both transforming as gauge fields, and one also explicitely verifies eq.(\ref{idiss}). Then, from eq.(\ref{ches}) we have \begin{equation} \bar \psi \gamma_\mu t^a \psi \to \pm \frac{i}{8\pi}\varepsilon_{\mu \nu \alpha} \partial_\nu A^a_\alpha \,\, . \label{Xsss} \end{equation} In writing eq.(\ref{Xsss}) we have ignored terms quadratic and cubic in the source which, as in $d=2$, are irrelevant for the current algebra. Correlation functions of currents pick a contribution from these terms, as already discussed in other approaches to bosonization \cite{dVR}-\cite{F}. Having these terms local support, they do not contribute to the current commutator algebra. (That this is so can be easily seen using for example the Bjorken-Johnson-Low method). We insist that our result (\ref{Xsss}) does not imply any kind of approximation. However, to achieve a complete bosonization, one needs an explicit local form for the bosonic action and it is at this point where approximations have to be envisaged so as to evaluate the fermion determinant. In $d=3$ dimensions this determinant cannot be computed exactly. However, all approximation approaches and regularization schemes have shown the occurrence of a parity violating Chern-Simons term together with parity conserving terms which can be computed approximately. We shall use the result obtained by making an expansion in inverse powers of the fermion mass \cite{Red}, \begin{equation} \ln \det (\id + m + /\kern-.52em b) = \pm \frac{i}{16\pi} S_{CS}[b] + I_{PC}[b] + O(\partial^2/m^2) , \label{9f} \end{equation} where the Chern-Simons action $S_{CS}$ is given by \begin{equation} S_{CS}[b] = \int\!d^3\!x\, \varepsilon_{\mu\nu\lambda} \mathop{\rm tr}\nolimits \int\!d^3\!x\, ( f_{\mu \nu} b_{\lambda} - \frac{2}{3} b_{\mu}b_{\nu}b_{\lambda} ) . \end{equation} Concerning the parity conserving contributions, one has \begin{equation} I_{PC}[b] = - \frac{1}{24\pi m} \mathop{\rm tr}\nolimits\int\!d^3\!x\, f^{\mu\nu} f_{\mu\nu} + \cdots , \label{8f} \end{equation} We can then write, up to corrections of order $1/m$, the bosonic action $S_{bos}[A]$ in the form (From here on we shall omit to indicate the trace {\it tr} for notation simplicity) \begin{eqnarray} \exp(-S_{bos}[A]) & = & \int {\cal{D}} b_\mu \, X[b] \exp(\pm \frac{i}{16\pi} S_{CS}[b]) \times \nonumber \\ & & \det(2\varepsilon_{\mu \nu \alpha}D_\nu[b]) \exp\left( \pm \frac{i}{16\pi} \int d^3x A_\mu {^*\!\!f}_\mu[b] \right) \label{ufin} \end{eqnarray} We shall now introduce ghost fields ${\bar c}_\alpha$ and $c_\alpha$ to write the determinant in the r.h.s. of eq.(\ref{ufin}). With this, $Z_{fer}[s]$ takes the form \begin{eqnarray} Z_{fer}[s] & = & X[s]^{-1} \int {\cal{D}} b_\mu {\cal{D}} {\bar c}_\alpha {\cal{D}} c_\alpha {\cal{D}} A_\mu \exp \left( \mp\frac{i}{16\pi} \int d^3x A_\mu {^*\!\!f}_\mu[s] \right) \times \nonumber \\ & & \exp(-S_{eff}[b,A,\bar c, c]) \label{312} \end{eqnarray} with \begin{eqnarray} S_{eff}[b,A,\bar c, c] & = & \mp \frac{i}{16\pi} S[b] \nonumber \\ & & \mp \frac{i}{8\pi} \varepsilon_{\mu \nu \alpha} \int d^3x ( A_\mu (\partial_\nu b_\alpha + b_\nu b_\alpha) - {\bar c}_\mu D_\nu [b] c_\alpha ) \label{312S} \end{eqnarray} and \begin{equation} S[b] = 2 \varepsilon_{\mu \nu \alpha} \int d^3x b_\mu(\partial_\nu b_\alpha +\frac{1}{3} b_\nu b_\alpha) \label{sssx} \end{equation} At this point we have arrived to an exact bosonization recipe for the fermion current, eq.(\ref{Xsss}), but we still need an explicit formula for the bosonic action as a functional of $A_\mu$. This requires integration over the auxiliary fields $b_\mu$, $\bar c_\mu$ and $c_\mu$ of the complicated effective action $S_{eff}$ as defined by eq.(\ref{312S}). In the two-dimensional case, this last step was possible because Polyakov-Wiegmann identity allowed us to decouple the auxiliary fields from the bosonic field ($A_\mu$). In the present case, integration will be possible because of the existence of an underlying BRST invariance that can be made apparent in $S_{eff}$. In order to directly get an {\it off-shell} nilpotent set of BRST transformations leaving invariant the effective action, we shall introduce additional auxiliary fields \cite{bas}, thus writing \begin{eqnarray} Z_{fer}[s] & = & X[s]^{-1} \int {\cal{D}} b_\mu {\cal{D}} {\bar c}_\alpha {\cal{D}} c_\alpha {\cal{D}} A_\mu {\cal{D}} h_\mu {\cal{D}} l {\cal{D}} \bar \chi \nonumber \\ & & \exp \left( \mp\frac{i}{16\pi} \int d^3x A_\mu {^*\!\!f}_\mu[s] \right) \exp(-{\tilde S}_{eff}[b,A,\bar c, c, h, l, \bar \chi]) \label{barra} \end{eqnarray} with $\tilde S_{eff}$ defined as \begin{eqnarray} & & {\tilde S}_{eff}[b,A,\bar c, c, h, l, \chi] = \mp \frac{i}{16\pi} S[b-h] \mp \frac{i}{16\pi} \int d^3x (l h_\mu h_\mu - 2 \bar \chi h_\mu c_\mu ) \nonumber \\ & & \mp \frac{i}{8\pi} \varepsilon_{\mu \nu \alpha} \int d^3x ( A_\mu (\partial_\nu b_\alpha + b_\nu b_\alpha) - {\bar c}_\mu D_\nu [b] c_\alpha ) \label{312SS} \end{eqnarray} Integration over the auxiliary field $l$ makes $h_\mu = 0$ this showing the equivalence of eq.(\ref{barra}) and eq.(\ref{312}). Now, the effective action ${\tilde S}_{eff}$ is invariant under BRST transformations defined as \[ \delta {\bar c}_\alpha = A_\alpha \;\;\;\;\;\; \delta A_\alpha = 0 \] \[ \delta b_\alpha = c_\alpha \;\;\;\;\;\; \delta c_\alpha = 0 \] \begin{equation} \delta h_\alpha = c_\alpha \;\;\;\; \delta \bar \chi = l \;\;\;\; \delta l = 0 \label{trb} \end{equation} This BRST transformations are related to those employed in the smooth bosonization \cite{DNS1}-\cite{DNS3}, \cite{DNS4}-\cite{DNS5} approach and resemblant of those arising in topological field theories \cite{BS}-\cite{BRT}. For example, in $d=4$ topological Yang-Mills theory the invariance of the starting classical action (the Chern-Pon\-trya\-gin topological charge) under the most general transformation of the gauge field, $b_\mu \to b_\mu + \epsilon_\mu$, leads to a BRST transformation for $b_\mu$ of the form $\delta b_\mu = c_\mu$, which corresponds to that in formula (\ref{trb}) \cite{Wi0}-\cite{BS}. Closer to our model are the so-called Schwartz type topological theories which include the Chern-Simons theory and the BF model analyzed in detail in refs.\cite{hor}-\cite{BRT}. It should be stressed that the topological character of the effective action (\ref{312S}) exclusively concerns the large fermion mass regime where the fermion determinant can be written in terms of the CS action. Now, using transformations (\ref{trb}), $\tilde S_{eff}$ can be compactly written in the form \begin{equation} \tilde S_{eff}[b,A,\bar c, c] = \mp \frac{i}{16\pi} S[b-h] \mp \frac{i}{8\pi} \int d^3x \, \delta{\cal F}[\bar c,b,h,\bar \chi] \label{sss} \end{equation} with \begin{equation} {\cal F} = \varepsilon_{\mu \nu \alpha} \bar c_\mu(\partial_\nu b_\alpha + b_\nu b_\alpha) + \frac{1}{2} \bar \chi h_\mu h_\mu \label{off} \end{equation} At this point, an arbitrary functional ${\cal G}$ may be added to ${\cal F}$ without changing the partition function since it will enter in $Z_{fer}$ as an exact BRST form. The idea is to choose ${\cal G}$ so as to decouple the auxiliary field $b_\mu$ (to be integrated out afterwards) from the vector field $A_\mu$ which will be the bosonic counterpart of the original fermion field. We shall then consider \begin{equation} {\cal F} \to {\cal F} + {\cal G} \label{s1} \end{equation} with \begin{eqnarray} {\cal G} & = & \frac{1}{2} \varepsilon _{\mu \nu \alpha }\bar c_\mu \; ([b_\nu,A_\alpha] + [A_\nu,A_\alpha] +C[b_\nu,h_\alpha] +(1+C)[A_\nu,h_\alpha] \nonumber \\ & & - (C+1) [h_\nu,h_\alpha] +2\partial_\nu b_\alpha +4\partial_\nu A_\alpha +2C \partial_\nu h_\alpha ) \label{el1} \end{eqnarray} Here $C$ is an arbitrary constant. The addition of $\delta {\cal G}$ allows us to make contact at this point with the effective action discussed in refs.\cite{B},\cite{vN}. Indeed, after the shift \begin{equation} b_\mu \to 2b_\mu - A_\mu + h_\mu \label{LG} \end{equation} (the new $b_\mu$ transforms again as a gauge connection, with $h_\mu$ transforming covariantly) the Lagrange multiplier $A_\mu$ (which will play the role of the bosonic field in our bosonization approach, as identified by the source term) completely decouples for $h_\mu = 0$, so that integrating out auxiliary fields we end with \begin{equation} Z_{fer}[s] = {\cal N}X[s]^{-1} \int {\cal{D}} A_\mu \exp ( \mp\frac{i}{16\pi} \int d^3x A_\mu {^*\!\!f}_\mu[s] ) \times \exp(\pm \frac{i}{16\pi}S_{CS}[A]). \label{chesi} \end{equation} Here ${\cal N}$ is a constant (i.e. it is independent of the source) resulting from integration of auxiliary, ghosts and the $b$ field, \begin{equation} {\cal N} = \int {\cal{D}} b_\mu \,\, det\left( 2(2+C)\varepsilon_{\mu\nu\alpha} D_\nu[b] \right) \exp(\pm \frac{i}{4\pi} S_{CS}[b]) \label{final} \end{equation} We have then the bosonization result \begin{equation} Z_{fer} [s] \approx Z_{CS}[s] \label{uli} \end{equation} where $\approx$ means that our result is valid up to $1/m$ corrections since we used a result for the fermion determinant which is valid up to this order. We then see that we have ended with a Chern-Simons action as the bosonic equivalent of the original free fermion action with a coupling to the external source $s_\mu$ of the form $ A_\mu {^*\!\!f}_\mu[s] $. In considering fermion current bosonization within the $1/m$ approximation, the following facts should be taken into account. It is at the lowest order in $1/m$ that the resulting bosonic action is topological and a large BRST invariance is unraveled. Now, using the freedom to modify the action by BRST exact forms, one could think of adding to the topological bosonic action terms of the form $ \delta{\cal H}$ with \begin{equation} {\cal H} = \int d^3x \ \varepsilon_{\mu \nu \alpha} {\bar c}_{\mu} {\cal H}_{\nu \alpha}[s] \, , \label{ad1} \end{equation} with ${\cal H}_{\mu \nu}[s]$ an arbitrary functional of the external source $s_{\mu}$. In particular, choosing adequately ${\cal H}$ one could think of changing or even eliminating, to this order in $1/m$, the source dependence from $Z_{fer}[s]$. Now, this is a characteristic of Schwarz like topological models \cite{BRT}. In particular, the phase space of the Chern-Simons theory is the moduli space of flat connections on the given space manifold. So, if one looks {\it up to this order in $1/m$}, at the generating functional of current Green functions, one has, from Eqs.(\ref{41}) and (\ref{9f}) that the generating functional of connected Green functions is precisely a Chern-Simons action for the source $s$, \begin{equation} W[s]=- log Z[s] = \mp \frac{i}{16\pi} S_{CS}[s]. \label{ad3} \end{equation} Thus, making functional derivatives in the above expression with respect to the source and then putting the sources to zero, all current vacuum expectation values vanish identically up to contact terms (these terms, derivatives of the Dirac delta function, also vanish if we regularize appropriately the product of operators at coincident points. Moreover our results are valid in the $m\to\infty$ limit where the deep ultraviolet region is excluded). Non-vanishing observables are in fact topological objets, non local functionals of $A_\mu$ (Wilson loops) that are in correspondance to knots polynomial invariants. Hence the bosonization recipe (\ref{Xsss}) when used to this order in $1/m$ makes sense if one is to calculate vacuum expectation values of fermion objects leading for example to holonomies in terms of the bosonic field $A_\mu$. This calculation was discussed at length in \cite{B}. \section{Summary} We have shown in this paper that the path-integral bosonization approach developed in previous investigations \cite{FS}-\cite{B} is well-suited to study fermion models in $d \ge 2$ dimensions when a non-Abelian symmetry is present. We have started by reobtaining in Section 2 the well-honoured non-Abelian bosonization recipe for two dimensional massless fermions. Although well-known, this result allowed us to identify the point in which the non-Abelian character of the symmetry makes difficult the factorization of the path-integral which will represent the partition function of the resulting bosonic model. In two dimensions this factorization can be seen as a result of the existence of the Polyakov-Wiegmann identity for Wess-Zumino-Witten actions, and this can {\it a priori} put some doubts on the possibility of extending the approach to $d>2$. That also in $d=3$ one can obtain very simple bosonization rules for the non-Abelian case is the main result of section 3. Concerning the fermion current, we obtained an exact bosonization result which is the natural extension of the Abelian case. In respect with the bosonization recipe for the fermion action, we considered the case of very massive fermions for which the fermion determinant is related to the non-Abelian Chern-Simons action. In this case the factorization of the auxiliary and Lagrange multiplier fields is achieved after discovering a BRST invariance reminiscent of that at the root of topological models and related to that exploited in the smooth bosonization approach \cite{DNS1}-\cite{DNS3}, \cite{DNS4}-\cite{DNS5}. Addition of BRST exact terms allows us to extract the partition function for the boson counterpart of the original fermion fields. Our bosonization method starts by introducing in the fermionic generating functional an auxiliary field as it is done in the smooth bosonization and duality approaches to bosonization \cite{DNS1}-\cite{mar}. It becomes clear in our approach that, for non-Abelian symmetries, it is crucial to include the ``Faddeev-Popov'' like determinant which accompanies the delta function imposing a condition on the auxiliary field curvature. In fact, the BRST symmetry which allowed to arrive to the correct bosonic generating functional can be seen as a result of this fact and related to the way in which BRST symmetry can be unraveled by a change of variables as advocated in ref.\cite{bas}. It should be stressed that the only approximation in our approach stems from the necessity of evaluating the fermion determinant which, in $d > 2$, implies some kind of expansion. In the present work we have used a result valid for very massive fermions but one can envisage approximations which can cover other regimes, in particular the massless case. This was considered for the abelian case in \cite{BFO} and the corresponding bosonization analysis thoroughly discussed in \cite {LNS}. We expect that a similar analysis can be done in the non-abelian case and we hope to report on it in a future paper. \newpage \section*{Appendix} \subsection*{ $d=2$} We shall prove identity (\ref{idle}) used in our derivation of $d=2$ bosonization rules, \begin{equation} \int {\cal{D}} b_\mu {\cal H}[b] \delta\left[\varepsilon_{\mu \nu} (f_{\mu\nu}[b] - f_{\mu\nu}[s])\right] =\int {\cal{D}} g {\cal H}[s^g] \label{idlex} \end{equation} where ${\cal H}$ is a gauge-invariant functional. Note that in eq.(\ref{idlex}) it is implicit that $b_\mu$ should be treated as a gauge field and hence a gauge fixing is required. A convenient gauge choice is \begin{equation} b_+ = s_+ \label{otrav} \end{equation} so that identity (\ref{idlex}) takes the form \begin{equation} \int {\cal{D}} b_+ {\cal{D}} b_- \;\Delta \delta(b_+ - s_+){\cal H}[b_+,b_-] \delta[\varepsilon_{\mu \nu} (f_{\mu\nu}[b] - f_{\mu\nu}[s]] = \int {\cal{D}} g {\cal H}[s^g] \label{idleg} \end{equation} with $\Delta$ the Faddeev-Popov determinant for gauge condition (\ref{otrav}), \begin{equation} \Delta = det D^{Adj}_+[s_+] \label{FP} \end{equation} We now prove eq.(\ref{idleg}). Let us start from the l.h.s. of eq.(\ref{idleg}) performing first the $b_+$ trivial integration and then the $b_-$ one \begin{eqnarray} & & \int {\cal{D}} b_+ {\cal{D}} b_- \Delta \delta(b_+ - s_+){\cal H}[b_+,b_-] \delta(\varepsilon_{\mu \nu} (f_{\mu\nu}[b] - f_{\mu\nu}[s]) = \nonumber\\ & & \Delta[s_+] \int {\cal{D}} b_- {\cal H}[s_+,b_-] \delta( D_+[s_+] b_- - D_+[s_+]s_- ) = \nonumber\\ & & \frac{\Delta[s_+]}{det D^{Adj}_+[s_+]} \int {\cal{D}} b_- {\cal H}[s_+,b_-] \delta(b_- - s_-) = {\cal H}[s_+,s_-] \label{largui} \end{eqnarray} In the last line we have used the explicit form of the Faddeev-Jacobian to cancel out both determinants. Being ${\cal H}[s_+,s_-]$ gauge independent, we can rewrite (\ref{largui}) in the form (appart from a gauge group volume factor) \begin{equation} \int {\cal{D}} b_+ {\cal{D}} b_- \Delta \delta(b_+ - s_+){\cal H}[b_+,b_-] \delta \left[\varepsilon_{\mu \nu} (f_{\mu\nu}[b] - f_{\mu\nu}[s]) \right] = \int {\cal{D}} g {\cal H}[s_+^g,s_-^g] \label{uno} \end{equation} Identity (\ref{idlex}) is then proven. \subsection*{ $d=3$} The proof of identity (\ref{IX}), the analogous in the $d=3$ case of (\ref{idlex}), is very simple. One wishes to prove that the generating functional $Z_{fer}[s]$ in the presence of a source $s_\mu$, \begin{equation} Z_{fer}[s] = \det (\id + m + /\kern-.52em s) = F[s] \label{leg16} \end{equation} can be written in the form \begin{equation} Z_{fer}[s] = {X[s]}^{-1} \int {\cal{D}} b_\mu \, X[b]F[b] \, \det( 2\varepsilon_{\mu \nu \alpha} D_\nu[b] ) \, \delta\left( \varepsilon_{\mu \nu \alpha} (f_{\nu \alpha}[b] - f_{\nu \alpha}[s])\right) \label{leg17} \end{equation} Here $X[b]$ is an arbitrary functional of $b_\mu$ satisfying $X[0^g]=1$. As advocated in \cite{FAS}, its introduction allows to end with a model in which the bosonic field transforms as a connection, this being consistent with the fact its dynamics is governed by a Chern-Simons action. The proof of eq.(\ref{leg17}) is based on the well-known identity \begin{equation} \delta(H[b]) = [\det (\delta H/\delta b)]^{-1} \delta(b - b^*) \end{equation} with $H[b^*] = 0$ and the fact that the equation $f_{\mu\nu}[b] = f_{\mu\nu}[s]$ has the unique solution $ b_\mu = s_\mu$. Let us end by noting that if one compares formula (\ref{leg17}) in $d=3$ dimensions with the corresponding one in $d=2$ (for example the identity (\ref{idlex})), one sees that a determinant equivalent to that appearing in the former is absent in the latter. This is due to the fact that the curvature condition requires three delta functions in $d=3$ dimensions but only one in $d=2$. Handling these delta functions leaves a jacobian in three dimensions while no jacobian remains in two dimensions. \vspace{1 cm} \underline{Acknowledgements}: The authors wish to thank Matthias Blau, Daniel Cabra, Fran\c{c}ois Delduc, C\'esar Fosco, Eric Ragoucy and Frank Thuillier for helpful discussions and comments. F.A.S. and C.N. are partially suported by Fundacion Antorchas, Argentina and a Commission of the European Communities contract No:C11*-CT93-0315.
proofpile-arXiv_065-416
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} Since the discovery of high temperature superconductors (HTSC), the pairing mechanism and the symmetry of the order parameter in these compounds are key questions at stake \cite{Dyn94,Sch94}. There are several experimental techniques which are able to address this problem. The experiments on quasiparticle tunneling \cite{Val91}, the linear temperature dependence of the penetration depth \cite{Har93}, the NMR and NQR measurements \cite{Kit93,Sli93}, and angular resolved photoemission experiments in Bi$_2$Sr$_2$CaCu$_2$O$_8$ \cite{She93,Ma} have yielded results consistent with $d$-wave pairing. On the other hand, quasiparticle tunneling, the exponential temperature dependence of the penetration depth, as well as the measurements of the electronic Raman scattering in Nd$_{2-x}$Ce$_x$CuO$_4$ are consistent with s-wave pairing \cite{Tra91,Mao94,Hac}. The measurements of the magnetic field dependence of the dc-SQUID (YBaCuO-Au-Pb arrangement) \cite{Woh94} clearly indicated d-wave behavior, while the experiments on single Josephson-junction Pb-YBCO \cite{Sun94} showed s-type behavior. So while the experimental evidence in favor of d-wave symmetry of the order parameter continuously grows, there is still no final consensus about it. Raman scattering is a powerful tool to address the problem of the symmetry of the order parameter. It allows to probe the symmetry of the scattering tensor by simply choosing different polarization directions of the incident and scattered light. From the investigations of the Raman scattering in conventional superconductors it is known that the superconducting transition manifests itself in a renormalization of the electronic Raman scattering intensity below T$_c$. It was found for Nb$_3$Sn and V$_3$Si \cite{Hac83,Die83} that normalized Raman spectra of these compounds show for temperatures below T$_c$ a peak associated with the pair breaking process at the energy 2$\Delta$, together with a strong decrease of the scattering intensity at frequencies lower than 2$\Delta$. In high temperature superconductors, the first measurements of electronic Raman scattering were reported in Refs.\onlinecite{Lyo87,Baz87,Hac88,Coo88}. But in this case the behavior of the electronic scattering differs from that in conventional superconductors: A pair breaking peak develops in the spectra below T$_c$, but the scattering intensity at frequencies below 2$\Delta$ does not show the usual sharp decrease. Instead, a monotonic decrease toward zero frequency is found. Moreover, for different symmetry components (A$_{1g}$, B$_{1g}$ and B$_{2g}$) the renormalization of the scattering intensity for T$<$T$_c$ is different and they exhibit peaks at different frequencies \cite{Hac88,Coo88,Hof94,Sta92,Hof95,Nem93,Che92,Che94,Dev94,Dev95,Kra94,Kra95}. These facts have been explained by Devereaux et al. \cite{Dev94,Dev95} in terms of d$_{x^2-y^2}$-wave pairing. Their calculations of the scattering cross-section have been performed for a cylindrical single-sheeted Fermi surface in the framework of the kinetic equation approach. The symmetry of the crystal was taken into account through calculating the Raman vertex, which was expanded in terms of a complete set of crystal harmonics defined on the Fermi surface. It was found that nontrivial coupling between the Raman vertex and an assumed strongly anisotropic energy gap leads to the strong symmetry dependence of the scattering intensity. The calculations \cite{Dev94,Dev95} predict specific symmetry dependencies of the low frequency scattering as well as the peak positions for the different symmetry components of the electronic Raman scattering at temperatures below T$_c$. The A$_{1g}$ peak position is sensitive to the parameters of the model calculation. It will appear below B$_{1g}$ peak position while with some parameters it may also appear at the B$_{1g}$ peak position. Nevertheless there is one set of parameters which can perfectly reproduce the experimental data\cite{Priv}. Later on this model was criticized by Krantz and Cardona \cite{Kra94,Kra95}. Their calculations \cite{Kra95} are based on the general description of the Raman scattering cross-section through the inverse effective mass tensor. In case of the multisheeted Fermi surface,(e.g. several CuO$_2$ planes per unit cell in HTSC) polarization dependent Raman efficiencies are determined by the averages of the corresponding effective mass fluctuations. The authors of Ref. \onlinecite{Kra95} used the effective masses from LDA band structure calculations for YBa$_2$Cu$_3$O$_7$ to determine the Raman scattering cross-section. They found that it contradicts the experimental results if one uses only d-wave pairing for a multisheeted Fermi surface of YBa$_2$Cu$_3$O$_7$. An explanation was given by assuming different types of the order parameter on different sheets of the Fermi surface. For a single-sheeted Fermi surface (i.e. one CuO$_2$-plane per unit cell) no mass fluctuations should occur. Therefore in the framework of the effective mass fluctuation approach, the A$_{1g}$ scattering component will be nearly totally screened and should peak at the same position where the B$_{1g}$ scattering component (2$\Delta_{max}$). Therefore straihgtforward measurements of the electronic Raman scattering in single-CuO$_2$ layered high-temperature superconductors (Tl-2201, La-214, Bi-2201, (Nd,Ce)-214) should clarifay this controversial point. Tl-2201 has the highest T$_c$ (up to 110K) \cite{Kol92} among the above mentioned single-CuO$_2$ layered compounds. Therefore all effects due to the gap opening are expected in the range of 300-600cm$^{-1}$, and they should not be obscured due to the Rayleigh scattering at small wavenumbers. Nevertheless, Raman measurements in only one pure scattering geometry (B$_{1g}$) are known \cite{Nem93,Blu94} for this compound, which showed \cite{Nem93} becides a T$_c$=80K two additional transitions at 100 and 125K, which can be indicative of the Tl-2212 and Tl-2223 phases. These facts lead us to reinvestigate the electronic Raman scattering in the Tl-based high temperature superconductor Tl-2201 (with different oxygen content) with one CuO$_2$-plane per unit cell. The comparison with the results of electronic Raman scattering experiments reported for the high temperature superconductors with several CuO$_2$-planes should clarify whether the multiband scattering is indeed important. We should mention that similar experiments on the single layered compound (La-214) were already carried out \cite{Che94}. Nevertheless, in the framework of comparison of the compounds with different number of CuO$_2$-planes the measurements on Tl-2201 are more favourable due to its high T$_c$. \section{EXPERIMENT} The investigated single crystals of Tl$_2$Ba$_2$CuO$_{6+\delta}$ (Tl-2201) had the shape of rectangular platelets with the size of approximately 2x2x0.2 mm$^3$. The two crystals investigated were characterized by a SQUID magnetometer. T$_c$ was found to be 90$\pm$3 K and 80$\pm$5 K. The crystals are slightly underdoped. It is known \cite{Shi92} that differences in T$_c$ in Tl-2201 compound originate from different oxygen concentrations. These crystals can be over- as well as underdoped. The heavily oxygen doped crystals show a metallic type of conductivity \cite{Shi92} and do not show a superconducting transition. The orientation of the tetragonal crystals was controlled by X-ray diffraction. Raman measurements were performed on "as grown" surfaces of the freshly prepared crystals. This is very important, because the crystal surface of Tl-based superconductors as well as of all high temperature superconductors is very sensitive to long exposure to air and especially to humid atmosphere. For the Raman measurements a DILOR XY triple spectrometer combined with a nitrogen cooled CCD detector was used. All Raman data were obtained at nearly backscattering geometry. The photon excitation was provided by the 488-nm line of an Ar$^+$ ion laser with laser power equal to 15W/cm$^2$. The estimated additional heating did not exceed 5K. \section{EXPERIMENTAL RESULTS} All measurements were performed with the polarization of the incident and scattered light parallel to the basal plane of the crystal, i.e. the CuO$_2$-planes. It was possible to measure the A$_{1g}$, B$_{1g}$, and B$_{2g}$ symmetry components of the Raman scattering cross-section. In addition to the previously published phonon peaks ($\approx$123cm$^{-1}$, $\approx$169cm$^{-1}$, $\approx$490cm$^{-1}$, $\approx590$cm$^{-1}$, $\approx$610cm$^{-1}$) \cite{Gas89,Kall94} we have detected some additional phonons ($\approx$240cm$^{-1}$, $\approx$300cm$^{-1}$, $\approx$330cm$^{-1}$, $\approx375$cm$^{-1}$) which we believe are the defect induced infrared active phonons. For all scattering geometries the spectra for temperatures well below T$_c$ were divided by the spectra just above T$_c$ in order to emphasize the redistribution of the scattering intensity in the superconducting state compared to the normal state. The results of the electronic Raman scattering in the crystals of Tl-2201 (T$_c$=80K,90K) are shown on Figs. \ref{f1}-\ref{f5}. In the crystal with T$_c$=80K the B$_{1g}$ scattering component measured in X'Y' configuration shows a well-defined peak at 430$\pm$15 cm$^{-1}$ (Fig. 1a). The X' and Y' axes are rotated by 45$^\circ$ with respect to the crystal X and Y axes, respectively, which are parallel to the crystallographic axes. The B$_{2g}$ scattering component in Fig. \ref{f1}b is less intense, but shows also a broad maximum with an average frequency of 380$\pm$35cm$^{-1}$. Raman spectra in the X'X' and XX geometries are presented in Fig. \ref{f2}a and b, showing spectra of A$_{1g}$ + B$_{2g}$ and A$_{1g}$ + B$_{1g}$ scattering components, respectively. In order to evaluate the A$_{1g}$ scattering component we subtracted the B$_{1g}$ and B$_{2g}$ components (see Fig. \ref{f1}a,b) from the XX and X'X' spectra, respectively. As one can see from Fig. \ref{f3}a,b the A$_{1g}$ scattering component peaks in both cases, at 345$\pm$20cm$^{-1}$. For the crystal with T$_c$=90K we found peaks of the B$_{1g}$, A$_{1g}$ and B$_{2g}$ scattering components at 460$\pm$15cm$^{-1}$, 350$\pm$20cm$^{-1}$, and 400$\pm$35cm$^{-1}$, respectively (Fig. \ref{f4}, lower panel). Another very important observation is that the low frequency behavior of the electronic Raman scattering exhibits strong anisotropy with respect to the symmetry components. One can see in Fig.4a (upper and lower panel) that the intensity decrease of the B$_{1g}$ scattering component toward lower frequencies fits the $\omega^3$ law predicted by Devereaux et al. \cite{Dev95}. For the A$_{1g}$ and B$_{2g}$ scattering components in Fig. \ref{f4}b and c, respectively, there is a linear decrease, which also agrees with the predictions by Deveraux et al. \cite{Dev95}. A summary of our results on Tl-2201 is presented in Table \ref{t1}. In order to follow the temperature behavior of the superconductor gap, we have measured the temperature dependence of the electronic Raman scattering. Following Devereaux et. al.\cite{Dev94,Dev95}, we assume that the peak in the B$_{1g}$ component of the electronic scattering corresponds to the value of 2$\Delta_{max}$. In Fig. \ref{f5}a and b, respectively, we show the B$_{1g}$ and A$_{1g}$ scattering component of Tl-2201 (T$_c$=80K) at different temperatures between 10K and T$_c$ divided by the spectrum at 100K. The experiments for the 90-K crystal yielded a similar behavior. With increasing temperature the intensity of the peak in Fig. \ref{f5}a associated with the pair breaking process decreases and the maximum shifts slightly to lower frequencies. Obviously, the temperature dependence of the superconductor gap does not follow the BCS behavior. In other words, upon cooling below T$_c$ the gap opens more abruptly than predicted by BCS theory. These results are similar to results reported for underdoped Bi-2212 \cite{Hof95}. Because the peak position of the A$_{1g}$ scattering component in Fig. \ref{f5}b has larger error bars compared to the B$_{1g}$ component, one cannot definitely say whether the data fit the BCS behavior or not. We also searched for superconductivity-induced changes in frequency and linewidth of the optical phonons. With the resolution of 1cm$^{-1}$ we have not observed such changes. Upon heating from 10K up to 200K the frequencies of all phonons decreased and the linewidths increased monotonically. \section {DISCUSSION} The Raman scattering intensity can be written in terms of the differential scattering cross section\cite{Dev95}: \begin{equation} \frac{\partial^2\sigma}{\partial\omega\partial\Omega}= \frac{\omega_s}{\omega_i}r_0^2S_{\gamma\gamma}(\vec{q},\omega) \end{equation} with \begin{equation} S_{\gamma\gamma}(\vec{q},\omega)=-\frac{1}{\pi}\left[1+n(\omega)\right]\Im m\chi_{\gamma\gamma}(\vec{q},\omega) \end{equation} Here r$_0=e^2/mc^2$ is the Thomson radius, $\omega_i (\omega_s)$ is the frequency of incident (scattered) photon, $\hbar$ and k$_B$ were set to 1. S$_{\gamma\gamma}$ is the generalized structure function, which is connected to the imaginary part of the Raman response function $\chi_{\gamma\gamma}$ through the fluctuation-dissipation theorem; $n(\omega)=1/[\exp (\omega/T)-1]$ is the Bose-Einstein distribution function. The Raman response function can be written as \cite{Kle84}: \begin{equation}\label{rares} \chi_{\gamma\gamma}(\vec{q},\omega)=\langle\gamma^2_{\vec{k}}\lambda_{\vec{k}} \rangle -\frac{\langle\gamma_{\vec{k}}\lambda_{\vec{k}}\rangle^2} {\langle\lambda_{\vec{k}}\rangle} \end{equation} with the Raman vertex $\gamma_{\vec{k}}$ written as \begin{equation} \gamma_{\vec{k}}(\omega_i,\omega_s)=\sum_L\gamma_L(\omega_i,\omega_s) \Phi_L(\vec{k}), \end{equation} where $\Phi_L(\vec{k})$ are either Brillouin zone or Fermi surface harmonics \cite{Dev95} which transform according to point group transformations of the crystal and $\lambda_{\vec{k}}$ is the Tsuneto function: \begin{equation} \lambda_{\vec{k}}\propto\frac{\left|\Delta_{\vec{k}}\right|^2} {\omega\sqrt{\omega^2-4\left|\Delta_{\vec{k}}\right|^2}}. \end{equation} The brackets $\langle\cdots\rangle$ in Eq. \ref{rares} denote an average of the momentum $\vec{k}$ over the Fermi surface. As is obvious, Raman scattering probes only $\left|\Delta\right|^2$. Therfore it is not possible to determine whether the gap function changes sign for different directions of $\vec{k}=(k_x,k_y)$ or not. But nevertheless the symmetry of the gap function can be inferred from the specific spectral features of each symmetry component of the electronic Raman scattering. For the gap of d-wave symmetry ($\Delta_{\vec{k}}=\Delta_{max}\cdot\cos 2\phi$, where $\phi$ is an angle between $\vec{k}$ and the a-axis), calculations \cite{Dev94,Dev95} predict different low frequency behavior for the different symmetry components. For B$_{2g}$ and A$_{1g}$ scattering components it should show a linear dependence in $\omega$, but for B$_{1g}$ it should be $\sim\omega^3$. The appearance of a power law in the low frequency scattering characterizes an energy gap which vanishes on lines on the Fermi surface. An appearance of $\omega^3$ law in B$_{1g}$ scattering component is specific for d$_{x^2-y^2}$-wave pairing \cite{Dev95}. The predicted values of the peak maxima for the B$_{1g}$, B$_{2g}$ and A$_{1g}$ scattering components are $\sim 2\Delta_{max}$, $\sim 1.6\Delta_{max}$ and $\sim 1.2\Delta_{max}$, respectively. These above mentioned peculiarities appear in our data. Indeed, the low frequency behavior of the B$_{1g}$ scattering component definitely differs from a linear behavior as seen in Fig. \ref{f4}a, whereas for the A$_{1g}$ and B$_{2g}$ scattering components it is linear in $\omega$ (see Fig. \ref{f4}b,c). For both crystals, the B$_{1g}$ scattering component peaks at a higher frequency than the B$_{2g}$, which in turn peaks at a higher frequency than the A$_{1g}$ component. Since Raman scattering does not probe the phase of the order parameter it is important to take into consideration other types of the pairing which can also have nodes on the Fermi surface, but do not change the sign, i.e. $s+id$-pairing, or strongly anisotropic s-pairing. For the $s+id$-pairing \cite{Dev95} $(\Delta (k)=\Delta_s + i\Delta_d\cos 2\phi)$ one gets the threshold at $\omega = 2\Delta_s$ (minimum pair breaking energy). While A$_{1g}$ and B$_{2g}$ scattering components exhibit a jump at this frequency, the B$_{1g}$ scattering component shows a continuous rise from zero and up to the peak at $\omega = 2\Delta_{max} =2\sqrt{(\Delta_s^2+\Delta_d^2)}$. The A$_{1g}$ and B$_{2g}$ scattering components also show broad maxima as in the case of pure d$_{x^2-y^2}$-wave pairing, but these maxima will be cut-off toward lower frequencies due to the strong jump at 2$\Delta_s$. Thus one should observe a low-frequency cutt-off in both A$_{1g}$ and B$_{2g}$ scattering components, which, however, is not observed in our data. For anisotropic s-pairing, showing the minimum of the gap on the diagonals of the two-dimensional Brillouin zone, ($\Delta (k) = \Delta_0 + \Delta_1\cos^42\phi$) one gets a single threshold on 2$\Delta_0$ for all scattering components as well as a peak at $\omega = 2\Delta_{max} = 2(\Delta_0 + \Delta_1)$ for the B$_{1g}$ scattering component. Therefore we will expect a picture which is very similar to the case of $s+id$-pairing, with one exception. The B$_{1g}$ scattering component should show an additional shoulder at the same position where the A$_{1g}$ and B$_{2g}$ scattering components show peaks \cite{Dev95}. This is also not the case for our data. In principle one can assume $\Delta_0$ to be very small or even zero. In this case one gets peaks at 2$\Delta_{max}$, 0.6$\Delta_{max}$ and 0.2$\Delta_{max}$ for the B$_{1g}$, B$_{2g}$ and A$_{1g}$ scattering components \cite{Dev95}, respectively. In addition, the low frequency behavior of the B$_{1g}$ scattering component will be linear. This also contradicts our results. Recently the model calculations of Devereaux et al. were criticized by Krantz and Cardona \cite{Kra94,Kra95}. The main argument against this theoretical model is that the realistic electronic band structure of the crystal is important, but that the one-sheeted Fermi surface approximation used by Devereaux et al.\cite{Dev95} is inappropriate. The authors of Ref.\onlinecite{Kra95} used a numerical model based on the LDA band structure calculations for YBaCuO in order to take into account the multisheeted Fermi surface of the superconductors with several CuO$_2$ -planes. It was pointed out that for the $\Delta = \Delta_0 cos 2\phi$ (d-wave pairing) and a multisheeted Fermi surface the calculations lead to a contradiction with the experiment, i.e. the A$_{1g}$ and B$_{1g}$ scattering components peak at the same position 2$\Delta_0$. In order to get consistency with the experiment, different types of the order parameter on different sheets of the Fermi surface were proposed. Only in this case the calculations in Ref.\onlinecite{Kra95} were able to get different positions of the maxima of the B$_{1g}$, A$_{1g}$, and B$_{2g}$ scattering components. For a one-sheeted Fermi surface the authors of Ref.\onlinecite{Kra95} found identical positions of the maxima for the A$_{1g}$ and B$_{1g}$ components, but a different position for the B$_{2g}$ component. Hence it was concluded that any difference in peak position of the A$_{1g}$ and B$_{1g}$ component is only consistent with multiband scattering of a multisheeted Fermi surface and different gap symmetries for each of the sheets. For superconductors with one CuO$_2$-plane, a multisheeted Fermi surface is invoked originating from Tl-like s-states \cite{Kra95} (Tl-2201) or from Sr-doping\cite{Kra} (La$_{2-x}$Sr$_x$CuO$_4$) in order to yield a difference in peak position for the A$_{1g}$ and B$_{1g}$ components. However, no experimental proof for such a Fermi surface contribution exists so far. Moreover, the calculations in Ref.\onlinecite{Kra95} failed in explaining of the symmetry-dependent low-frequency dependence of the Raman scattering intensity, whereas this important experimental fact was observed not only in our experiments, but also in Bi-Sr-Ca-Cu-O \cite{Sta92,Hof95,Dev94}, Y-Ba-Cu-O\cite{Hac88,Coo88} and La-Sr-Cu-O \cite{Che94} systems. In addition, it is obvious that all superconductors with different crystal structures have a different electronic structure. Hence, if the multiband scattering model would be crucial we would expect absolutely different behavior for the different superconductors which is actually in contradiction with existing experimental results. Even if one compares the superconductors with the same number of CuO$_2$ planes, one finds that, while the interplanar distance (distance between Cu atoms in different CuO$_2$-planes) is quite similar, the dimpling (in-plane Cu-O-Cu angle) differs very much from compound to compound (see Table \ref{t2}). YBa$_2$Cu$_3$O$_7$ exhibits the largest dimpling compared to other compounds. Strong dimpling should lift the degeneracy of otherwise identical CuO$_2$-planes or Fermi surface sheets. This dimpling can strongly affect the LDA calculations because the interplanar interaction should depend on this parameter. And finally on top of that, use of the effective mass approach is very much questionable in case of high temperature superconductors, because this approach can be used only for nonresonant Raman scattering \cite{Abr}. In high temperature superconductors we, however, are always in the regime of the resonant scattering. Moreover, the electron correlation effects in HTSC are not treated sufficiently by LDA. In contrast to the conclusion of Ref. \onlinecite{Kra95} our experiments show that the one-CuO$_2$-plane compound Tl-2201 shows very similar behavior compared to compounds with several CuO$_2$-planes, such as Tl-2223, Bi-2212 and YBaCuO \cite{Hac88,Coo88,Hof94,Sta92,Hof95,Nem93,Che92,Che94,Dev94,Dev95,Kra94,Kra95}. We also found that the frequency of the B$_{1g}$ maximum scales with T$_c$, and it corresponds to the value 2$\Delta_{max}/k_BT_c=7.6\pm 0.4$. This value is very close to the values (with the exception of Nd$_{2-x}$Ce$_x$CuO$_4$ \cite{Hac}) found for other high temperature superconductors as shown in Table \ref{t3}. The temperature dependence of the gap (B$_{1g}$ component in Fig. \ref{f5}a) in our experiment differs from the BCS behavior, i.e. upon cooling the gap opens more abruptly than predicted by BCS theory. This is consistent with the spin fluctuation theory of high temperature superconductivity \cite{Mon92}, favoring d$_{x^2-y^2}$-wave pairing. The model considers pair binding as well as pair breaking effects due to the spin fluctuations. Gap opening leads to a suppression of low-frequency spin fluctuations and therefore to reduced pair-breaking. Therefore in underdoped crystals (we consider our Tl-2201 crystals as underdoped) this effect will lead to a more abrupt opening of the gap upon cooling below T$_c$ compared to BCS behavior. In conclusion, we presented measurements of the electronic Raman scattering on high-T$_c$ Tl-2201 single crystals with one CuO$_2$-plane per unit cell. The peculiarities of the electronic Raman scattering, i.e. the power law frequency dependence of the diferent scattering components at low frequencies, their different peak positions as well as the values of 2$\Delta_{max}/k_BT_c=7.6\pm 0.4$ are found to be very similar in compounds with one and several CuO$_2$ planes. All nearly optimally doped high-T$_c$ superconductors (with the exception of (Nd,Ce)-214 \cite{Hac}) show a very similar behavior of the electronic Raman scattering consistent with the d$_{x^2-y^2}$- wave symmetry of the underlying order parameter. \section {ACKNOWLEDGMENTS} This work was supported by DFG through SFB 341 and by BMBF FKZ 13 N 6586. One of the authors L.V.G acknowledges support from the Alexander von Humboldt Foundation and expresses his gratitude for the hospitality at the 2.Physikalisches Institut RWTH-Aachen.
proofpile-arXiv_065-417
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} \label{intro} Granular fluids are dense media composed of elementary elements of macroscopic size, undergoing collisions in which their {\it macroscopic} energy is not conserved. In the last decade, the flow of this granular fluids has received a great attention from the physics community, both because these media offer a "simple" example of dissipative systems and because of their numerous industrial applications. Two factors make the behaviour of granular fluids very different from that of molecular fluids. Firstly, the macroscopic size of the particles implies that external fields (boundaries or gravity) have a much stronger effect on granular fluids. Secondly, the energy of the granular fluid is not a conserved variable, since the heat dissipated in collisions can be considered as lost as far as the flow is concerned. These two effects are often difficult to disentangle in experiments, and also in the numerical simulations that aim at a realistic modelling of these experiments \cite{HERRMAN}. A different type of numerical simulations was initiated by several groups \cite{Goldhirsch, Young1,Young3,Young2,Bernu}. In this approach external fields are ignored, and only the dissipation is taken into account. This dissipation is moreover modelled in a very simplistic way, by describing the particles as monodisperse rigid hard spheres undergoing inelastic collisions. The dissipation is entirely specified by the restitution coefficient $r$, where $1-r$ is the fraction of the kinetic energy lost in a collision (in the center of mass frame). The aim of these simulations is not to model actual experiments involving granular fluids. Instead, the models are used to assess the difference in behaviour induced by the dissipation between a granular fluid and its "atomic" counterpart ($r=1$, the hard sphere fluid.) In particular, the simulations can be used to investigate the validity of the hydrodynamic equations frequently used to describe granular flow in more realistic situations \cite{Haff,Campbell}. As for atomic fluids, they also provide a direct way of measuring the equation of state and transport coefficients that enter these equations. These transport coefficients can then be compared to those obtained using "granular kinetic theory" \cite{Jenkins}. A particularly simple and instructive situation that can easily be studied in numerical simulations is the 2-dimensional "cooling problem" first studied in \cite{Goldhirsch, Young2}. In their simulations, these authors start from an equilibrium configuration of an {\it elastic} ($r=1$) hard disc fluid. The behaviour of this fluid after introducing a nonzero restitution coefficient is followed using the standard molecular dynamics method for hard bodies \cite{AT87}, with a collision rule that takes the dissipation into account. The kinetic temperature (average kinetic energy per particle) decreases due to the inelasticity of the collisions, so that the fluid cools down as time increases. It was shown in \cite{Goldhirsch, Young2} that, depending on the system size, on the restitution coefficient and density, this cooling follows different routes. In the simplest case, (small systems or small dissipation) the fluid remains homogeneous at all temperatures. In larger systems or for larger dissipations, either the velocity field or the density field in the fluid develop instabilities and become inhomogeneous. It was also found by the same authors that the occurence of such instabilities is in qualitative agreement with the predictions of a linear stability analysis of the hydrodynamic equations for granular fluids \cite{Haff}. Finally, it was discovered in \cite{Young1,Young2} that in some cases, the cooling ends at a finite time due to a singularity in the system dynamics, which was shown to correspond to an infinite number of collisions within a finite time. This singularity, observed both in 1 and 2 dimensions, was described as an "inelastic collapse" of the system. In this paper, a detailed and quantitative analysis of the instability of homogeneous cooling of granular fluids is attempted. Our aim is to compare quantitatively the predictions of granular kinetic theory and granular hydrodynamics to the results of molecular dynamics simulations of the cooling of an inelastic hard disk fluid. The paper is organized as follows. The main predictions of granular hydrodynamics concerning the cooling problem are briefly recalled. Computational details concerning the simulation are given in section 3. The different regimes occuring during the cooling are analyzed in section 4, and compared with the theoretical predictions. Our main focus will be on the growth rate of the density instability, that can be computed by monitoring the structure factor of the system as a function of time. Finally, the problem of the "inelastic collapse" is adressed in section 5, where a possible method for avoiding this singularity in the system dynamics is proposed. \section{Hydrodynamic analysis of the cooling problem} The hydrodynamic equations that have been proposed to describe granular flow \cite{Haff} are based on mass and momentum conservation, and are very similar to the usual Navier-Stokes equations. The only modification is the appearance of a new term in the energy (or temperature) equation, accounting for the loss of energy in the collisions. These equations can be compactly written in the form \begin{eqnarray} \Dhyd{\ensuremath {\rho}}{t} & = & -\ensuremath {\rho} \Div{ \ensuremath {\Vect{v}}} \\ \ensuremath {\rho} \Dhyd{\ensuremath {\Vect{v}}}{t} & = & -\Div{\Tens{P}} \\ \ensuremath {\rho} \Dhyd{T}{t} & = & -\Div{\Vect{Q}}-tr \left( \Tens{P}\,\Tens{D} \right )-\gamma T^\frac{3}{2} \end{eqnarray} where $D/Dt$ is the hydrodynamic derivative, $\Tens{D} $ the symmetrized velocity gradient tensor, \Tens{P} the stress tensor, \Vect{Q} the heat flux and $\gamma$ represented the rate of energy lost due to inelastic collisions. For a hard disk fluid, the equation of state and the expression of the various transport coefficients can be obtained from Jenkins and Richman kinetic theory \cite{Jenkins}. These expressions are recalled in Appendix A. The energy sink term, $\gamma T^{3/2}$, has also been written in the form appropriate for hard disks. $\gamma$ in that case is a function of the density and the restitution coefficient, which at least in the low density limit must be proportionnal to $ \ensuremath {\rho} $ and $(1-r)$. This can be understood from the following reasoning: the kinetic energy loss per particle per unit time is proportional to the collision frequency (i.e. to $\ensuremath {\rho} T^{1/2}$) and to the energy loss per collision $(1-r) T$. A trivial solution of the cooling problem formulated above corresponds to an homogeneously cooling fluid, with a uniform density, a vanishing velocity field, and a uniform temperature with an algebraic time decay \begin{eqnarray} \label{temperature} T(t)=T_0{\left(1+\frac{t}{t_0}\right)}^{-2} \end{eqnarray}. Here $t_0=2 \ensuremath {\rho}_0/(\gamma_0 {T_0}^{1/2})$ sets the time scale for temperature decay in the fluid. The linear stability of this homogeneous solution has been investigated in references \cite{Goldhirsch, McNamara}. For completness, the main steps of this analysis will be repeated here. The linearized equations that describe the evolution of a sinusoidal perturbation \begin{eqnarray} \ensuremath {\delta \rho}=\ensuremath {{\delta \rho}_{\Vect{k}}} \exp \left ( i \ensuremath {\Vect{k}} \cdot r \right ) \\ \ensuremath {\delta \V}=\ensuremath {{\delta \V}_{\Vect{k}}} \exp \left ( i \ensuremath {\Vect{k}} \cdot r \right ) \\ \ensuremath {\delta T}=\ensuremath {{\delta T}_{\Vect{k}}} \exp \left ( i \ensuremath {\Vect{k}} \cdot r \right ) \end{eqnarray} around the homogeneous solution are \begin{eqnarray} \label{cont1} \Der{\ensuremath {{\delta \rho}_{\Vect{k}}}}{t} & = & -i\ensuremath {\rho}_0 \left ( \ensuremath {\Vect{k}} \cdot \ensuremath {{\delta \V}_{\Vect{k}}} \right ) \\ \label{eq:vitesseParral2} \ensuremath {\rho}_0 \Der{\ensuremath {\Vect{k}} \cdot \ensuremath {{\delta \V}_{\Vect{k}}}}{t} & = & -ik^2 \left [ \ensuremath {\rho}_0 p'( \nu_0)\ensuremath {{\delta T}_{\Vect{k}}}+T_0 \left (p'(\nu_0)+\nu_0 \left ( \Der {p'}{\nu} \right )_0 \right )\ensuremath {{\delta \rho}_{\Vect{k}}} \right ] \nonumber\\ & & \mbox{ } - \mu_0 \left [ k^2 \left ( \ensuremath {\Vect{k}} \cdot \ensuremath {{\delta \V}_{\Vect{k}}} \right ) \right ] \\ \label{eq:vitessePerp2} \ensuremath {\rho}_0 \Der{\ensuremath {\Vect{k}}_\perp \cdot \ensuremath {{\delta \V}_{\Vect{k}}}}{t} & = & - \mu_0 \left [ k^2 \left ( \ensuremath {\Vect{k}}_\perp \cdot \ensuremath {{\delta \V}_{\Vect{k}}} \right ) \right ] \\ \label{eq:temperature2} \left [ \ensuremath {{\delta \rho}_{\Vect{k}}} \Der{T_0}{t}+\ensuremath {\rho}_0 \Der{\ensuremath {{\delta T}_{\Vect{k}}}}{t} \right ] & = & -\kappa_0 k^2 \ensuremath {{\delta T}_{\Vect{k}}} -{p_h}_0\left ( i \ensuremath {\Vect{k}} \cdot \ensuremath {{\delta \V}_{\Vect{k}}} \right) \nonumber\\ & & \mbox{ } - \frac{3}{2} \gamma_0 {T_0}^\frac{1}{2} \ensuremath {{\delta T}_{\Vect{k}}} - T^{\frac{3}{2}} \nu {\left( \Der{\gamma}{\nu} \right )}_0 \frac{\ensuremath {{\delta \rho}_{\Vect{k}}}}{\ensuremath {\rho}_0} \end{eqnarray} As for usual fluids, the transverse part of the velocity field completely decouples from the longitudinal part, and decays with time as $ \left ( 1 + {t}/{t_0} \right )^{-k^2 T_0^{1/2} {t_0}/{\ensuremath {\rho}_0}} $. The longitudinal part of the velocity field, the temperature and the density are coupled, and give rise to three modes that have an algebraic time dependance \begin {eqnarray} \ensuremath {{\delta \rho}_{\Vect{k}}} & = & \ensuremath { \delta \widetilde{\rho}}_{\Vect{k}} { \left [ 1+ t/t_0 \right ]}^\ensuremath {\xi} \\ \ensuremath {{\delta \V}_{\Vect{k}}} & = & \ensuremath { \delta \widetilde{\V}}_{\Vect{k}} { \left [ 1+ t/t_0 \right ]}^{\ensuremath {\xi}-1} \\ \ensuremath {{\delta T}_{\Vect{k}}} & = & \ensuremath { \delta \widetilde{T}}_{\Vect{k}} { \left [ 1+ t/t_0 \right ]}^{\ensuremath {\xi}-2} \end{eqnarray}. The exponents $\xi(k)$ for the three modes are the three roots of the determinant of the following set of equations \begin{eqnarray*} \ensuremath { \delta \widetilde{\rho}}_{\Vect{k}} \left (\frac{\ensuremath {\xi}(\ensuremath {\xi}-1)}{t_0^2}+ \ensuremath {\Vect{k}}^2 T_0 \left [p'(\nu_0)+\nu_0 \left ( \Der {p'}{\nu} \right )_0 \right ] + \frac{\mu_0 k^2} {\ensuremath {\rho}_0 t_0} \ensuremath {\xi} \right ) + \ensuremath { \delta \widetilde{T}}_{\Vect{k}} \left [ k^2 \ensuremath {\rho}_0 p'(\nu_0) \right ] & = & 0 \\ \ensuremath { \delta \widetilde{\rho}}_{\Vect{k}} \left [ \frac{\ensuremath {T}_0}{t_0} \left (-2 + T_0^{\frac{1}{2}} \nu {\left( \Der{\gamma}{\nu}\right )}_0 \frac{t_0}{\ensuremath {\rho}_0} - p'(\nu_0) \ensuremath {\xi} \right ) \right ] + \ensuremath { \delta \widetilde{T}}_{\Vect{k}} \left [ (\ensuremath {\xi}+1) \frac{\ensuremath {\rho}_0}{t_0} + \kappa_0 k^2 \right ] & = & 0 \end{eqnarray*} A typical plot of the wavevector dependance of these three roots, together with the growth rate of the velocity perturbations, is shown in figure 1. It must be emphasized that the stability of velocity disturbances is determined by the comparison between the growth exponent of the disturbance with the value $-1$ that characterizes the decay of the {\it thermal} velocity. Hence a growth exponent larger than $-1$ for the transverse or longitudinal velocity fields is indicative of an instability of the macroscopic velocity. If the growth exponent of the longitudinal velocity field is larger than $-1$, a corresponding instability in the density field will follow from equation \ref{cont1}. This analysis yields to the prediction of three different possible behaviours of the system, depending on the value of the parameters and on the system size, that introduces a lower wavevector cutoff. If this lower cutoff corresponds to the line $C$ of figure 1, the homogeneous solution will be linearly stable. This regime will be described as the homogeneous kinetic regime. If the lower cutoff moves to the abscissa indicated by line $B$ in figure 1, the transverse velocity field will become unstable while the system remains homogeneous. In this "shearing" regime, first observed in reference \cite{Goldhirsch}, a shearing flow will develop in the system. Finally, for a lower cutoff corresponding to abscissa $A$, an instability of the longitudinal velocity field and the corresponding instability in the density field will take place together with the shearing instability. In this "clustering" regime, the growth of density disturbances will yield to the formation of dense clusters of particles, as first observed in reference \cite{Hopkins}. All three situations have already been observed in numerical simulations of the cooling in two dimensional granular fluids. The aim of the next sections will be to attempt a quantitative analysis of the behaviour of a cooling granular fluid, and to compare the results to the predictions summarized above. \section{Computational details} The model simulated in this work is in all respects similar to that studied in \cite{Goldhirsch, Young2}. The system is made up of $N$ hard inelastic disks of diameter $\sigma$, in a square cell of size $L$ with periodic boundary conditions. The cell size $L$ sets the lower cutoff in wavevector space, $k_{min}=2\pi/L$. A standard cell-linked Molecular Dynamics algorithm for hard bodies \cite{AT87} is used. In a first step, the system is equilibrated with a coefficient $r$ equal to unity. At time $t=0$, inelasticity is switched on and cooling starts, with an initial temperature $T_0$. The restitution coefficient enters through a simple modification of the standard collision rule between hard disks, the velocities of the two disks after a collision being given by \begin{eqnarray} \label{coll1} \Vect {u_1}'=\Vect{u_1}-\frac{1}{2}(1+r) [\hat{\Vect{n}} \cdot (\Vect{ u_1}-\Vect {u_2}) ] \hat{\Vect{n}} \\ \label{coll2} \Vect {u_2}'=\Vect{u_2}+\frac{1}{2}(1+r) [\hat{\Vect{n}} \cdot (\Vect{ u_1}-\Vect {u_2}) ] \hat{\Vect{n}} \end{eqnarray} where the primes denote the quantities after collision and $ \hat{{\Vect{n}}} $ is a unit vector along the centers line from particle 1 towards particle 2. The natural units in this problem are the particle mass $m$ and diameter, and the thermal energy at $t=0$, i.e. $T_0$. The corresponding time unit is $\tau=(m/T)^{1/2}\sigma$. The state of the system is defined by three dimensionless numbers, which are the reduced size $L/\sigma$ (or equivalently the reduced cutoff $k_{min}^*= k_{min}\sigma$), the reduced density $\rho*=\sigma^2 N/L^2 $, and the restitution coefficient $r$. The state of the fluid during the cooling was monitored by a systematic computation of coarse-grained (hydrodynamic) density and velocity fields. The coarse graining is obtaining here from a division of the system into 100 square subcells. Besides, statistical quantities characterizing the state of the system have also been systematically computed. These quantities are the momenta of the velocity distribution of individual particles, the pair correlation function $g(r)$ for interparticle distance, and the structure factor \begin{eqnarray} S(\Vect{k}) & = & \frac{1}{N} \rho_\Vect{k} \rho_\Vect{-k} \end{eqnarray} This structure factor can be computed for all wavevectors compatible with the periodic boundary condition, of the form $(n_x,n_y) k_{min}$. As the system is not in a stationary state, these quantities are time dependant. A large enough system is thus necessary to obtain reasonable statistics without time averaging. The values of $N$ investigated in this work vary from $N=1600$ to $N=10000$. \section{results} \subsection{Kinetic regime} According to the analysis of section \ref{intro}, the kinetic regime corresponding to a stable homogeneous cooling will be observed (at a given density and restitution coefficient) for small enough systems. Such a situation allows a clear testing of some of the hypothesis of the kinetic theory description of the granular fluid. In particular, the pair correlation function and velocity distribution can be compared to that of an elastic hard disk fluid throughout the cooling process. The temperature decay can be monitored and compared to the theoretical prediction (equation \ref{temperature}), and the decay time $t_0$ (or equivalently the coefficient $\gamma(\rho)$) compared to the prediction of kinetic theory. The pair correlation of an homogeneously cooling granular fluid after the temperature has dropped by a factor of 10 is shown in figure 2. This comparison shows that the local structure of the cooling granular medium (which determines its equation of state) remains essentially identical to that of an equilibrium fluid. The study of the velocity distribution function shows that this distribution remains maxwellian throughout the cooling. This similarity between the structure and velocity distribution of the granular fluid and the usual hard disk fluid suggests that the kinetic theory of Jenkins \cite{Jenkins} is applicable. This expectation is borne out by the study of the time dependance of the fluid temperature. As shown in figure 3, the temperature decay is perfectly described by equation \ref{temperature}. The density dependance of the decay time $t_0$ is compared in figure 4 to the prediction of kinetic theory (see appendix B). The agreement is extremely good, and suggests that all the transport coefficients appearing in the hydrodynamic equations can be estimated using this kinetic theory. \subsection{Shearing regime} If the restitution coefficient $r$ decreases or if the size of the system increases, the hydrodynamic theory predicts a regime in which transverse fluctuations of the velocity field are unstable. This regime is indeed observed in the simulations, as shown in figure 5. A shear flow that corresponds to the smallest wavevector compatible with the periodic boundary conditions develops in the system. In this regime, the total kinetic energy of the system (which in that case is not the temperature, since the system has developped an ordered flow pattern) appreciably deviates from equation \ref{temperature}, as shown in figure 6. \subsection{Clustered regime} For even larger systems or smaller restitution coefficients, the cooling granular fluid becomes inhomogeneous, as shown in figure 7. this spontaneous formation of density inhomogeneities (or clusters) was first observed in the simulations of the cooling problem by Goldhirsch and Zanetti and Young and McNamara \cite{Goldhirsch, Young2}. Two different explanations have been put forward to explain this cluster formation. The first one, found in \cite{Goldhirsch}, is to consider this cluster formation as a secondary instability of the shearing regime, due to the developpment of temperature and pressure gradients in the shearing regime. The second possible explanation is that cluster formation is directly related to the linear instability of the density modes predicted by hydrodynamic theory. In order to characterize quantitatively this clustering regime, the structure factor $S(k,t)$ of the system has been computed as a function of time and wavevector. The corresponding data is shown in figure 8. The growth of the density inhomogeneities results in the appearance of a low wavevector peak in the structure factor, that rapidly increases with time. According to hydrodynamics, the time dependance of $S(k,t)$ should be algebraic, i.e. \begin{eqnarray} S(\ensuremath {\Vect{k}},t) & = & S(\ensuremath {\Vect{k}},0) {\left ( 1 + \frac{t}{t_0} \right )}^{2\xi(\ensuremath {\Vect{k}})} \end{eqnarray} so that the ratio \begin{eqnarray} \frac{\ln (S(\ensuremath {\Vect{k}},t))-\ln (S(\ensuremath {\Vect{k}},0))}{ln \left ( 1 + \frac{t}{t_0} \right )} & = & 2\xi(\ensuremath {\Vect{k}}) \end{eqnarray} should be independent of time. This ratio is plotted in figure 9 as a function of wavevector for different times. $2\xi(k)$ seems to be reasonably independent of time, and its low wavevector value appears to be consistent with the prediction of linearized hydrodynamics. Hence the density instability can be interpreted as resulting from a linear instability of the homogeneous solution of the hydrodynamic equations. Note that it was recently observed by McNamara and Young that the "clustering" fluid eventually develops for long times into an ordered flow pattern of the "shearing" type. This is also consistent with hydrodynamics, since the growth rate of the transverse velocity modes is positive. The description of the formation of this shearing flow in an inhomogeneous system, however, is beyond the possibilities of linearized hydrodynamics. \section{Inelastic collapse and how to avoid it} The inelastic collapse singularity was first observed by \cite{Bernu, Young1} in simulations of unidimensional inelastic system. This collapse can be described as the appearance of an infinite number of correlated collisions between a few particles, taking place in a finite time. The same phenomenon was observed in two dimensions by \cite{Young2}. It was shown that in that case the correlated collisions take place between a small number of essentially aligned particles, so that the unidimensional situation is practically reproduced. In order to avoid this inelastic collapse, a slightly modified collision rule between the particles can be introduced. At each collision, the relative velocity of the two particles is first computed according to the usual rule (equations \ref{coll1} and \ref{coll2}), then rotated by a small (less than 5 degrees) random angle. This can be justified by invoking the unavoidable roughness of actual solid particles, conservation of angular momentum being (virtually) ensured by a transfer to the internal degrees of freedom of the particles. As to inelastic collapse, the aim of this modified collision rule is to hinder the formation of correlated particle lines that cause this singularity. Indeed, inelastic collapse was not observed in the simulations where this "random" collision rule was used, while under the same conditions a system following the "deterministic" collision rule always underwent inelastic collapse (figure 10). Hence inelastic collapse appears to be a pathology related to the use of purely specular collision rule between particles, rather than a characteristic of inelastic fluids. \section{Conclusion and perspectives} The main objective of this work was to assess the validity of the hydrodynamic description of granular fluids originally proposed by \cite{Haff}, and of the kinetic theory calculation of the associated transport coefficients. The study of the particularly simple "cooling fluid" case and of the associated instabilities provides an ideal benchmark for this description. The comparison between numerical simulations and theoretical predictions in this simple case shows that the theory is quantitatively accurate. A similar conclusion was also reached in a recent study by McNamara and Young \cite{preprint}, who showed that the transitions between the different cooling regimes were correctly predicted by the theory. The description of the inelastic collapse phenomenon observed by McNamara and Young is obviously beyond the possibilities of kinetic theory or hydrodynamics. It was shown that this phenmenon can easily be avoided by introducing a small amount of randomness in the collisions between particles, similar to what would be caused by the natural roughness of granular particles. Obviously, a correct description of granular fluid cannot be achieved without a knowledge of the boundary conditions that must be used for the hydrodynamic equations. These conditions, and in particular those that correspond to the very important case of vibrating solid walls, are not known. Their determination, through the quantitative comparison of numerical simulation and theory, will be the subject of future work. \section*{Ackowledgments} This work was supported by the Pole Scientifique de Mod\'elisation Num\'erique at ENS-Lyon. \pagebreak
proofpile-arXiv_065-418
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{ Introduction} \label{Introduction} Relativistic wind models of Gamma Ray Bursts (GRB) are a recent development of the dissipative relativistic fireball scenario which is a natural consequence of observationally set requirements irrespective of the GRB distance scale (e.g.~\cite{me95}). The observed bursts are expected to be produced in optically thin shocks in the later stages of the fireball expansion. Two types of shocks have been considered, based on different interpretations of the burst duration and variability. First, the ``external'' deceleration shocks (\cite{rm92}; \cite{mr93}) that develop at $r_{dec}$, due to the unavoidable interaction of the relativistic ejecta with the surrounding medium, give bursts whose duration is determined (in the ``impulsive'' regime) by the relativistic dynamic timescale $t_{dec} \sim r_{dec}/c \Gamma^2$, where $\Gamma$ is the Lorentz factor of the expansion (e.g. \cite{mr93}; \cite{ka94}; Sari, Narayan \& Piran 1996). Second, the ``internal'' dissipation shocks in an unsteady relativistic wind outflow (\cite{rm94} [RM94]; \cite{paxu94}), where both the burst duration $t_w$ and time variability $t_{var} \leq t_w$, are characteristic of the progenitor mechanism (e.g. a disrupted disk accretion timescale, dynamic or turbulent timescales, etc.). Simple ``external'' shocks tend to produce fairly smooth light curves with a fast rise followed by an exponential decay (FRED), as observed in some bursts. However, many bursts have multi-peaked, irregular light curves; those may find a more natural explanation (Fenimore, Madras \& Nayakshin 1996) in the relativistic wind scenario mentioned above. While the spectral properties of impulsive ``external'' shocks have been studied in some detail (\cite{mrp94} [MRP94]; \cite{sa96}), so far, only rough estimates have been made for wind ``internal'' shock spectra (\cite{mr94}; \cite{th94}). It is of great interest to explore the spectral properties of both types of shock scenarios, particularly in view of the increasingly sophisticated analyses of both GRB light curves (e.g. \cite{fe96}; \cite{nor96}) and spectral characteristics (e.g. \cite{band93}). In this letter, we investigate the spectral properties of ``internal'' shocks for representative parameter values and discuss those properties in the light of anticipated results from HETE. \section{ Physical Model and Shock Parameters} \label{sec:model} In the unsteady relativistic wind model, the GRB event is caused by the release of an amount of energy $E=10^{51}E_{51}$ erg, inside a volume of typical dimension $r_{\ell}= 3 \times 10^{10} t_{var}$ cm, over a timescale $t_w$ (RM94). The average dimensionless entropy per particle $\eta= \langle E/(M c^2) \rangle - 1 = \langle L/({\dot M} c^2)\rangle$ determines the character of the overall flow. The bulk Lorentz factor of the flow increases as $\Gamma \propto r$ at first and later saturates to a value $\Gamma = \Gamma_s$ at $r \mathrel{\mathpalette\simov >} r_s \approx r_{\ell} \Gamma_s $, where $\Gamma_s \sim \eta $ for the values of $\eta$ that we consider here (eq. [\ref{eta_lim}]). It is assumed that $\eta$ varies substantially, $\Delta \eta \sim \eta$, on a timescale $t_{var} $. As an example, we consider shells of matter with two different $\eta$ values ejected at time intervals $t_{var}$. After saturation these will coast with bulk Lorentz factors $\Gamma_f \sim q_f \Gamma_s $ and $\Gamma_r \sim q_r \Gamma_s$, where $ q_r < q_f \sim 1$. Shells of different $\eta$ catch up with each other at a dissipation radius $r_{d} \approx \Gamma_{f} \Gamma_{r} c t_{var} \approx 3 \times 10^{14}\; q^2 \Gamma_{s_2}^{2} \; t_{var} \; $ cm (see RM94, eq. [4]), where $q =\sqrt{q_f q_r}$ and the subscript 2 (3) stands for quantities measured in units of $10^{2}$ ($10^3$). This gives rise to a $\em{forward}$ and a $\em{reverse}$ shock which are mirror images of each other in the contact discontinuity (center of momentum, CoM) frame. The CoM moves in the lab frame with $\Gamma_{b}\approx \sqrt{ \Gamma_{f} \Gamma_{r}}$, for $\Gamma_s \gg 1$. In the CoM frame the two shocks move in opposite directions with small Lorentz factors $\Gamma'_{sh} -1 \approx (1/2) \;[({\Gamma_{f}}/{\Gamma_{r}}) + ({\Gamma_{r}}/{\Gamma_{f}})]^{1/2}-1 \ge 10^{-2}$, for $\Gamma_{f}, \Gamma_{r} \gg 1$, while in the lab frame both shocks move forward with Lorentz factors $\Gamma_{f}$ and $ \Gamma_{r} $ respectively. The shocks compress the gas to a comoving frame ( primed quantities refer to the $\em{comoving}$ frame, which is the CoM frame ) density \begin{equation} n\acute{}\; (r_{d}) \approx \frac{8 \times 10^{9}}{q^4} \frac{E_{51}}{\theta ^{2} t_{w} t_{var}^{2}} \frac{1} { \eta_{2} \Gamma_{s_2}^{5} } \; \; \; \rm{cm^{-3}}, \end{equation} where $\theta^{2}$ is the normalized opening solid angle of the wind ($1$ if spherical). The bulk kinetic energy carried by the flow is randomized in the shocks with a mechanical efficiency $\varepsilon_{sh} \approx (q_f + q_r - 2 q)/ (q_f+ q_r)$ (eq. [1] in RM94). In order for the shocks to produce a non-thermal spectrum and radiate a substantial portion of the wind energy, the dissipation must occur beyond both the wind photosphere (see section 2.2 in RM94) and the saturation radius. This restricts $\eta$ to the range \begin{equation} 33 \; \left( \frac{E_{51}}{\theta^{2} t_{w} t_{var}} \right)^{1/5} \mathrel{\mathpalette\simov <} \eta \mathrel{\mathpalette\simov <} 80 \; \left( \frac{E_{51}}{\theta^{2} t_w t_{var}}\right)^{1/4}. \label{eta_lim} \end{equation} The collisionless ``internal'' shocks accelerate particles to relativistic energies with a nonthermal distribution. We parameterize the post-shock electron energy by a factor $\kappa$, so that the average random electron Lorentz factor is $\gamma_e \sim \kappa\Gamma\acute{}_{sh} \sim \kappa \le m_p/m_e $ (with the protons' being $\gamma_p \sim \Gamma\acute{}_{sh}$). An injection fraction $\zeta$ of the post-shock electrons is accelerated to a power-law distribution ($ d N(\gamma) \sim \gamma^{-p} d\gamma$, for $\gamma_{min} \leq \gamma $). We take $p \ge 3$, so that most of the energy is carried by the low energy part of the electron spectrum ($\gamma_{min} \approx ((p-2)/(p-1)) \kappa$). A measure of the efficiency of the transfer of energy between protons and electrons behind the shock is given by $\varepsilon_{pe} \simeq \zeta \kappa/(m_p/m_e) \sim \zeta\kappa_3$. Here, we will use $\zeta=1$ and $\kappa\le 10^3$ ( but see \cite{byme96}) and $p=3$ which is consistent with the fits to observed spectra (\cite{hanlon94}). The electrons get accelerated at the expense of the randomized proton energy behind the shocks, which, for an individual shock (of which there are $t_w/t_{var}$), represents a fraction $ ({t_{var}}/{t_w}) \varepsilon_{pe} \varepsilon_{sh} $ of the total energy $E_0$. Magnetic fields in the shocks can be due to a frozen-in component from the progenitor, or may build up by turbulence behind the shocks. We parameterize the strength of the magnetic field in the shocks by the fraction $\lambda$ of the magnetic energy density to the particle random energy density ($u\acute{}_{\scriptscriptstyle B} = \lambda n\acute{} m_p c^2$). The comoving magnetic field therefore is \begin{equation} B\acute{} \approx \frac{1.72\times 10^{4}}{ q^2 \Gamma_{s_2}^2} \; \sqrt{\frac{ E_{51}\varepsilon_{sh}} {\theta^{2} t_w t_{var}^{2}} \; \frac{\lambda} {\eta_2 \Gamma_{s_2}}} \quad \hbox{G}. \end{equation} The relativistic electrons will lose energy due to synchrotron radiation, and inverse Compton (IC) scattering of this radiation. The respective radiative efficiencies are $\varepsilon_{sy} = t_{sy}^{'-1}/(t_{sy}^{'-1} +t_{ic}^{'-1} +t_{ex}^{'-1})$, and $\varepsilon_{ic} = t_{ic}^{'-1}/(t_{sy}^{'-1} +t_{ic}^{'-1} +t_{ex}^{'-1}) $ . The timescales are defined below. The comoving synchrotron timescale is determined by the least energetic electrons : \begin{equation} t'_{sy} \approx 8 \times 10^8 /\gamma_{min} B^{'2} = 5.2 \; \displaystyle q^4 \; \frac{\theta^2 t_w t_{var}^2}{E_{51} \varepsilon_{sh}}\; \frac{\eta_2 \Gamma_{s_2}^5}{\lambda \kappa_{sy_3}} \; \; \; \hbox{ms}, \end{equation} where the subscript {\em sy} ({\em ic}) refers to the synchrotron (IC) emitting shock. The IC cooling depends on, and competes with, synchrotron cooling. IC cooling dominates if the magnetic field is weak, a large fraction of electrons are accelerated and share the protons' momentum very efficiently (i.e. $\lambda /(\zeta \kappa_{ic_3} )\ll 1$); while synchrotron cooling dominates in the opposite case. The IC timescale ($t'_{ic} \approx 3 \times 10^7/u\acute{}_{sy} \gamma_{min}$) for the two limiting cases is \begin{equation} t'_{ic} \approx \left\{ \begin{array}{ll} 9.6 \;\displaystyle \frac{q^4}{\zeta \varepsilon_{sh}} \frac{\theta^2 t_w t_{var}^2}{E_{51}} \frac{\eta_2 \Gamma_{s_2}^5}{<\gamma_3^2>} \frac{\gamma_{*_3}}{\kappa_{ic_3}} \left[1 +\frac{t\acute{}_{sy}(\gamma_*)}{t\acute{}_{ex}}\right] & \mbox{ms $\;\;\;$if IC dominates}\\ \label{t_ic_ic} \\ 4.95\; \displaystyle \;\frac{q^4}{\varepsilon_{sh}} \frac{\theta^2 t_w t_{var}^2}{E_{51}}\; \frac{\eta_2 \Gamma_{s_2}^5} {\sqrt{\zeta \lambda \kappa_{ic_3} <\gamma_3^2>}} & \mbox{ms \quad if synchr. dominates} \label{eq:t_ic_sy}. \end{array} \right. \end{equation} Here $\gamma_* $ is the electron Lorenz factor that corresponds to the peak emission frequency, $\gamma_* =max \left[\gamma_{min}, \gamma_{abs} \right]$, $\gamma_{abs}$ is defined below equation (\ref{v_abs}) and $ <\gamma^2_3> = 2\times 10^{-6} \displaystyle \gamma_{min}^2 \;ln(\gamma_{max}/{\gamma_*})$. The lab frame shell width (shocked region) is $\Delta r \sim r_d \Gamma_{b}^{-2} \sim \alpha c t_{var}$, and the comoving crossing time $t_{ex}' \approx 10^{2} (\alpha q^2) \; \Gamma_{s_2} \;t_{var}$ s provides an estimate for the adiabatic loss time. The spectrum of an ``internal'' shock burst consists of two synchrotron and four IC components (coming from all shocks' combinations). The synchrotron components are characterized by up to three {\it break frequencies}, given below in the lab frame: i) The $\gamma_{min}$ break frequency $\nu' = 8 \times 10^5 B' \gamma_{min}^{2}$, which in the lab frame gets blue-shifted by $\Gamma_{f}$ ($\Gamma_r$) for the forward (reverse) shock. Using $q_{sh}=[q_f,q_r]$ to refer to the shock, we have \begin{equation} h\nu_{min} \approx 1.45\; \displaystyle \;\frac{q_{sh}}{q^2} \; \frac{\kappa_{sy_3}^2}{ \Gamma_{s_2} } \;\sqrt{\frac{E_{51} \varepsilon_{sh}}{\theta^2 t_w t_{var}^2} \frac{\lambda}{\eta_2 \Gamma_{s_2}}} \quad \hbox{keV}. \label{v_sy_min} \end{equation} ii) The self absorption frequency $\nu_{abs}$. If the electron power law starts at low enough energies, the radiation field becomes optically thick at a frequency determined by $\frac{3}{2} m_{e} \gamma_{abs} \nu_{abs}^{'2} = F'_{\nu_{abs}}$, and is obtained by solving a non-linear algebraic equation; in the limit $\nu_{abs} \gg \nu_{min}$ it is \begin{equation} h \nu_{abs} \approx 18.4 \;\displaystyle \frac{q_{sh}}{q^{10/7}}\; \frac{1}{\Gamma_{s_2}} \left( \frac{\lambda}{\eta_2 \Gamma_{s_2}} \right)^{1/14} \left ( \frac{E_{51}\varepsilon_{sh} } {\theta^{2} t_{w} t_{var}^{2}} \right)^{5/14}\; (\zeta \varepsilon_{sy} \kappa_{sy_3})^{2/7} \mbox{ eV $\; \; \;$ for $\gamma_{min} < \gamma_{abs}$}, \label{v_abs} \end{equation} where $\gamma_{abs} = 9 \times 10^3 \sqrt{\nu_{abs}/(q_{sh} \Gamma_{s_2} B\acute{})}$. iii) The frequency where the photon spectrum of the minimum energy electrons becomes optically thick ($\nu_{\scriptscriptstyle{RJ}}$), and below which it assumes a Rayleigh-Jeans spectrum slope. This happens at $\nu_{min}$ when $\gamma_{min}< \gamma_{abs}$; if $\gamma_{min} \gg \gamma_{abs}$ it is \begin{equation} h\nu_{\scriptscriptstyle{RJ}} \approx 0.15 \; \displaystyle \frac{q_{sh}}{q^{4/5}} \left(\frac{E_{51} \varepsilon_{sh}} {\theta^{2} t_{w} t_{var}^{2}} \right)^{1/5} \left(\frac{\eta_2}{\lambda} \right)^{2/5} \left(\frac{\zeta \varepsilon_{sh}}{\Gamma_{s_2}}\right)^{3/5} \frac{1}{\kappa_{sy_3}^{8/5}} \; \; \hbox{eV}. \label{eq:v_RJ} \end{equation} iv) A cutoff is expected at $\nu_{max} = (\gamma_{max}/\gamma_{min})^{2} \; \nu_{min}$, where $\gamma_{max}$ is the electron energy in each shock at which the electron power law cuts off due to radiative or adiabatic losses. It is determined by $t_{acc}(\gamma_{max}) \approx t'_{cool}(\gamma_{max})$, where $t_{acc}$ is the electron acceleration timescale in the shocks (a multiple $10 \times A_{10}$ of the inverse gyro-synchrotron frequency), $t_{acc} \approx {3.57 \times 10^{-6} A_{10}}{B'(r_{d})^{-1}} \gamma$ s, and $t_{cool}$ is the minimum of all the radiation and adiabatic cooling timescales involved, $t'_{cool_{r,f}} = min \left\{ t_{ex}', t'_{sy_{f,r}}, t'_{IC_{r}, } t'_{IC_{f}} \right \}$ . Each synchrotron component can be characterized by three frequencies in ascending order, $\nu_{sy,j}^{i}$, where $i=1,..,3$ and $j=1$ for the reverse and $j=2$ for the forward shock. Similarly, the pure and combined IC spectra are characterized by the frequencies $\nu_{ic,j}^{i} \approx ({4}/{3}) \;\kappa_{ic}^{2} \nu_{sy,l}^{i}$ where, $j=1,..,4$ (1 corresponds to {\it IC reverse} , 2 to {\it IC forward}, 3 to {\it IC reverse-forward} and 4 to {\it IC forward-reverse}), and if j is odd, $l=1$, otherwise $l=2$. The shape of each component depends on the relationship between the relevant $\gamma_{min}, \gamma_{abs} $ and $\gamma_{max}$. In a power per logarithmic frequency interval plot, the fluence of each component exhibits a peak of $ S_{i} = 1.6 \times 10^{-6} (E_{51}/(\theta D_{28})^{2}) \; \varepsilon_{sh} \varepsilon_{i} \zeta \kappa_{i_3} \; \; {\rm erg/cm}^{2}$, where $i=1 (2)$ for synchrotron (IC), and $ D_{28}$ is the luminosity distance corresponding to $z \approx 1$ for a flat Universe, with $H_{o} \approx 80$ ($D_{L} = (2 c/H_{o}) (1+ z - \sqrt{1+z} \approx 3 \; \hbox{Gpc})$). The spectrum is obtained by adding up these six components. In practice, the forward and reverse components have values very close to each other and they essentially merge. The resultant spectrum is then checked for the effects of pair production; the $\gamma\gamma$ optical depth is calculated for each comoving frequency above $m_e c^2 /k$ using the number density of photons above the corresponding threshold, as obtained from the initial spectrum; finally the spectrum is modified accordingly. We note that most of the scattering in our spectra occurs in the Thomson regime, unlike in ~\cite{sa96}, who consider large $\gamma$ in the framework of impulsive shock models. ( Klein-Nishina corrections may become relevant in a few $\kappa \mathrel{\mathpalette\simov <} 10^2$ cases, but at frequencies which lie in the $\gamma\gamma$ absorbed part of the spectrum). \section{ Typical Wind Spectra} We discuss here the properties of some representative spectra. We assume a total event energy of $E=10^{51}$ erg and a geometry of a spherical section, or jet, of opening angle $\theta = 0.1$ (the physics is the same as in a spherical wind , provided $\theta > \Gamma^{-1}$). We have investigated a range of dynamic parameters ($\eta, t_w$ and $t_{var}$), and used $q_f=2, q_r=0.5$, and $\kappa_f =\kappa_r$. In figure {\ref{fig_long}} we present spectra for a long burst ($t_w =100 $ s) and in figure {\ref{fig_short}} for a short one ($t_w= 1$ s), for different $\eta, \lambda$ and $\kappa$ values. The sharper spectral features present would be smoothed by inhomogeneities in a real flow. The range of $\eta$ where a substantial fraction of the wind energy is radiated with nonthermal spectra is considerably restricted (eq. [\ref{eta_lim}]) and implies lower values than in impulsive models. Larger $\eta$ may be difficult to produce in a natural way, while lower values lead to shocks below the photosphere that make bursts too dim to be observed. The allowed volume of parameter space is not very large (\cite{ha96} [PM96]). Nonetheless, it allows for an appreciable variety of spectra, since models differing only slightly in $\eta$ can produce significantly different spectra. This is due to the strong dependence of the photon number density on $\Gamma_s$ (i.e. $ n\acute{}_{\gamma} \propto \Gamma_s^{-7} $) which determines pair production. Lower $\kappa$ spectra are less affected by pair production, since they contain less energetic electrons and therefore fewer photons. Generally, other parameters being equal, the higher $\eta$ models tend to produce spectra that span over a wider range of frequencies. For spectra like the majority of the ones observed by BATSE (\cite{Fi&Me_rev}) values of $\kappa < 10^2$ and $\lambda > 10^{-1}$ are excluded. Low $\kappa$ spectra are either too dim or too soft; high $\lambda$ brings up the synchrotron component and may be appropriate for a small percentage of bursts with low frequency excess (\cite{Preece96}). In general, for a given $\kappa$ value a wide range of $\lambda$ values (3 - 6 orders of magnitude) is allowed, the trend being that higher $\kappa$ values must combine with lower $\lambda$ values, fairly independently of $\eta$ (within the allowed range of values). This trend is due to the fact that lower magnetic fields require higher electron energies in order for the break to fall in the BATSE window (e.g. eq. [\ref{v_sy_min}], [\ref{v_abs}] ). High energy power laws (like those reported in {\cite{hanlon94}}) are common (see fig. \ref{fig_short} and right column of fig. \ref{fig_long}). For a more complete discussion see PM96. The effect of a longer $t_w$ is to push the pair cutoff to higher frequencies, because, for fixed $E_0$, the flux and photon density are lower. For the same $\eta$, $\kappa$ and $t_{var}$, longer bursts require higher $\lambda$. A greater $t_{var}$ with the rest of the parameters unchanged would again require higher $\lambda$ in order to produce observed-like spectra. The cases considered here were chosen with $ t_w < t_{dec} \approx (E_0/n_{ex} m_p c^5 \eta^8)^{1/3}$ s. \section{Conclusions} \label{conclusions} An unsteady relativistic wind provides an attractive scenario for the generation of GRB, since it requires smaller bulk Lorentz factors than the impulsive models, i.e. it can accommodate higher baryon loads (RM94). In addition, the lack of kinematic restrictions (e.g. \cite{fe96}) allows it, in principle, to describe events with arbitrarily complex light curves. We have calculated spectra from optical through TeV frequencies for bursts with a range of durations and variability timescales. At cosmological distances the total energies and photon densities implied by the model are likely to turn those spectra optically thick to $\gamma\gamma$ pair production. Most of them may therefore be missed by, or show a high energy cut-off in, the EGRET window, but they would be prominent in the BATSE and HETE gamma-ray windows. Low frequency (down to 5 keV) excess reported recently ({\cite{Preece96}}) may be attributed to a pronounced synchrotron component due to a relatively high magnetic field, in a fraction of the bursts. The continued propagation of an unsteady wind flow should eventually lead to its deceleration by the surrounding medium. If the latter is of any appreciable density, it could lead to another burst with the ``external'' shock characteristics (MRP94, \cite{mr94}), provided that not all of the wind energy was radiated away by the ``internal'' ones ($\varepsilon_{sh} < 1$). The spectra of the ``internal'' shocks are different from those coming from the ``external'' ones, the main differences being that the former cover a narrower range of frequencies and most have high energy cutoffs due to pair production opacity (PM96). If GRB progenitors have escaped their parent galaxies and are in a low density intergalactic medium ($n_{ex} <10^{-3}$ cm$^{-3}$), the ``external'' impulsive shocks would be of long duration ($> 3\times10^3 /\eta_2^{8/3} $ s) and low intensity, and most would be totally missed, hence the GRB could be entirely due to unsteady wind ``internal'' shocks. If the GRB occur in a denser medium (e.g. the galactic ISM), the last peak of multiple-peaked GRB would be smooth and FRED-like but the earlier peaks, due to unsteady wind ``internal'' shocks, may be arbitrarily complicated. Those might also be responsible for delayed GeV emission ({\cite{mr94}}). A number of unknown parameters enter into the calculation of GRB models, and HETE may provide the information required to narrow the range of their allowed values, as well as a test for the general scenario of unsteady relativistic winds. For the HETE sensitivities indicated in figures \ref{fig_long} and \ref{fig_short}, we expect that some bursts will show a simultaneous X-ray counterpart. However, detection by the Ultraviolet Transient Camera (UTC) on HETE should be rare for the majority of bursts (note though that figures \ref{fig_long} and \ref{fig_short} refer to $z\sim$ 1 distances, and that at a few $\times 10^2$ Mpc a bright burst would be $30-100$ times brighter, thus increasing the likelihood of detection by the UTC). For the majority of faint bursts, a detection in the ultraviolet would be possible only for low values of $\kappa$, resulting in a very broad and flat spectrum with upper cutoffs, if any, only at the highest energies. \acknowledgements{This research has been supported through NASA NAG5-2362, NAG5-2857 and NSF PHY94-07194. We are grateful to the Institute for Theoretical Physics, UCSB, for its hospitality, and to participants in the Nonthermal Gamma-Ray Source Workshop for discussions}. \pagebreak
proofpile-arXiv_065-419
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction and Sample Selection} \subsection{Background} The surface brightness fluctuation (SBF) method of distance determination works by measuring the ratio of the second and first moments of the stellar luminosity function in a galaxy. This ratio, called $\overline L$, is then the luminosity-weighted, average luminosity of a stellar population and is roughly equal to the luminosity of a single giant star. In terms of magnitudes, this quantity is represented as $\overline M$, the absolute ``fluctuation magnitude.'' What we measure, of course, is the apparent fluctuation magnitude in a particular photometric band, in our case the $I$~band, \ifmmode\overline{m}_I\else$\overline{m}_I$\fi. In order to be useful as a distance estimator, \ifmmode\overline{m}_I\else$\overline{m}_I$\fi\ must be calibrated, either empirically, by tying the measurements to the Cepheid distance scale, or theoretically, according to stellar population synthesis models. Tonry and Schneider (1988) were the first to quantify the SBF phenomenon. Their method was based on a measurement of the amount of power on the scale of the point spread function in the power spectrum of a CCD image. They applied this method to images of the galaxies M32 and NGC 3379. Subsequent work by Tonry, Luppino, and Schneider (1988) and Tonry, Ajhar, \& Luppino (1989, 1990) revised and refined the analysis techniques and presented further observations in $V$, $R$, and $I$ of early-type galaxies in Virgo, Leo, and the Local Group. Tonry et al.\ (1990) found that the $I$~band was most suitable for measuring distances and attempted to calibrate the SBF method theoretically using the Revised Yale Isochrones (Green, Demarque, \& King 1987). There were obvious problems with this calibration, however, so a completely empirical calibration for \ifmmode\overline{M}_I\else$\overline{M}_I$\fi\ was presented by Tonry (1991). The calibration was based on the variation of \ifmmode\overline{m}_I\else$\overline{m}_I$\fi\ with \ifmmode(V{-}I)_0\else$(V{-}I)_0$\fi\ color in the Fornax cluster and took its zero point from the Cepheid distance to M31. Tonry used this calibration to derive the Hubble constant. A detailed review of the modern SBF technique can be found in Jacoby et al.\ (1992), which also provides some historical context for the method. In recent years, the SBF technique has been used to measure distances and study a variety of stellar populations in several different bands. $K$-band SBF observations have been reported by Luppino \& Tonry (1993), Pahre \& Mould (1994), and Jensen, Luppino, \& Tonry (1996). These studies find that $\overline m_K$ is also a very good distance estimator. Dressler (1993) has measured $I$-band SBF in Centaurus ellipticals, finding evidence in support of the Great Attractor model. Lorenz et al.\ (1993) have measured $I$-band SBF in Fornax, and Simard \& Pritchet (1994) have reported distances to two Coma~I galaxies using $V$-band SBF observations. Ajhar \& Tonry (1994) reported measurements of \ifmmode\overline{m}_I\else$\overline{m}_I$\fi\ and $\overline m_V$ for 19 Milky Way globular clusters and considered the implications for both the distance scale and stellar populations. Tiede, Frogel, \& Terndrup (1995) measured \ifmmode\overline{m}_I\else$\overline{m}_I$\fi\ and $\overline m_V$ for the bulge of the Milky Way and derived the distance to the Galactic center. Sodemann \& Thomsen (1995, 1996) have used fluctuation colors and radial gradients to investigate stellar populations in galaxies. Finally, an enormous amount of progress has been achieved on the theoretical SBF front through the stellar population models of Worthey (1993a,b, 1994), Buzzoni (1993, 1995), and Yi (1996). \subsection{Genesis of the SBF Survey} When it became apparent that $I$-band SBF observations could indeed provide accurate and reliable distances to galaxies, we undertook a large survey of nearby galaxies. The sample selection is not precisely defined because the measurement of SBF depends on a number of criteria which are not ordinarily cataloged, such as dust content. In addition, because the measurement of SBF is fairly expensive in terms of telescope time and quality of seeing, it simply was not possible to observe all early-type galaxies within some magnitude limit out to a redshift which is large enough to make peculiar velocities negligible. Nevertheless, we have tried to manage fairly complete coverage of early-type galaxies within 2000~km/s and brighter than $B = 13.5$, and we have significant coverage beyond those limits. Comparison with the Third Reference Catalog of Bright Galaxies (de Vaucouleurs et al.\ 1991) reveals that of the early-type galaxies ($T<0$) with $B \le 13.5$ in the RC3, we have observed 76\% with heliocentric velocity $v < 1000$~km/s, 73\% with $1000 < v < 1400$, 64\% with $1400 < v < 2000$, 49\% with $2000 < v < 2800$, and we have data for more than 40 galaxies with $v > 2800$~km/s. Virtually all of the galaxies closer than $v<2000$ where we lack data are S0s for which measuring SBF is complicated by dust and/or disk/bulge problems, and since many of them are in the cores of clusters such as Virgo, we do not regard their distances as being important enough to delay completion of our survey. The survey is, however, an ongoing project, with some data still to be reduced. About 50\% of our sample is listed as E galaxies ($T\le-4$), about 40\% as S0s, and 10\% as ``spirals'' ($T\ge0$). Our sample of galaxies is drawn from the entire sky, and the completeness was mainly driven by the vicissitudes of weather and telescope time, so the sampling is fairly random. The survey includes a large number of galaxies in the vicinity of the Virgo supercluster, and the next paper in this series will present an analysis of their peculiar motions. The following section describes the SBF survey in more detail, including the observations, photometric reductions, and consistency checks. In Section 3 we use our observations of galaxies in groups to derive the dependence of \ifmmode\overline{M}_I\else$\overline{M}_I$\fi\ on \ifmmode(V{-}I)_0\else$(V{-}I)_0$\fi. Seven of these groups also have Cepheid distances, which we use to set the zero point of the \ifmmode\overline{M}_I\else$\overline{M}_I$\fi--\ifmmode(V{-}I)\else$(V{-}I)$\fi\ relation. This new \ifmmode\overline{M}_I\else$\overline{M}_I$\fi\ calibration agrees well with theory and supersedes the old calibration of Tonry (1991). We then compare our distances to those found using a number of other methods. In Section 4 we discuss the tie to the large-scale Hubble flow and implications for the value of $H_0$. The final section provides a summary of our main conclusions. \section{Observations and Reductions} All in all, the SBF survey extends over some 40 observing runs at 6 telescopes. Table 1 lists these runs along with some salient information. Note that the date of the run is coded in the name as (Observatory)MMYY; the remaining columns are described below. The normal observing procedure when the skies were clear was to obtain sky flats each night and observe a number of Landolt standard stars. We preferred observing the faint standard star fields of Landolt (1992) in which there are several stars per CCD field and where the observations are long enough that shutter timing is not a problem. Table 2 gives our usual fields. During a typical night we would observe about 10 fields comprising perhaps 50 stars over a range of airmasses from 1.1 to 2.0. We also strived to observe stars over a wide range of color ranging from $0 < \ifmmode(V{-}I)\else$(V{-}I)$\fi < 2$. Because there is substantial fringing seen in the $I$ band with thin CCDs, at some point in a run we would spend several hours looking at a blank field in order to build up a ``fringe frame''. We have found the fringing pattern for a given CCD and filter (although not the amplitude) to be remarkably stable from night to night (even year to year). Hence, a single fringe frame was used to correct an entire run's data, and we usually collected a new one for each run. The reductions of the photometry proceed by bias subtraction, flattening, and following Landolt (1992), summing the net flux from photometric standards within a 14\arcsec\ aperture. We also estimate a flux error from the sky brightness and variability over the image and remove any stars whose expected error is greater than 0.02 magnitude. Once all the photometric observations from a run have been reduced, we fit the results according to \begin{equation} m = m_1 - 2.5\log(f) - A\;\sec z + C\;\ifmmode(V{-}I)\else$(V{-}I)$\fi, \end{equation} where $f$ is the flux from the star in terms of electrons per second. We have found that $m_1$ and $C$ are constant during a run with a given CCD and filter, so we fit for a single value for these parameters and extinction coefficients $A$ for each night. The rms residual of the fit is typically about 0.01 magnitude which is satisfactory accuracy for our purposes. Table 1 lists typical values for $m_1$, $A$, and $C$ for each run in the two filters $V$ and $I$. Note the havoc in the extinction caused by the eruption by Pinatubo in 1991. Galaxy reductions proceed by first bias subtraction, division by a flat field, and subtraction of any fringing present in $I$ band data. We always take multiple images of a galaxy with the telescope moved by several arcseconds between images, and determine these offsets to the nearest pixel. Any bad pixels or columns are masked out, and the images are shifted into registration. We next run a program called ``autoclean'' which identifies cosmic rays in the stack of images and removes them by replacement with appropriately scaled data from the rest of the stack. Autoclean also gives us an estimate of how photometric the sequence of observations was by producing accurate flux ratios between the exposures. Finally we make a mask of the obvious stars and companion galaxies in the cleaned image and determine the sky background by fitting the outer parts of the galaxy image with an $r^{1/4}$ profile plus sky level. This is usually done simultaneously for $V$ and $I$ images, and when the sky levels are determined, we also compute \ifmmode(V{-}I)_0\else$(V{-}I)_0$\fi\ colors as a function of radius from the center of the galaxy. In order to knit all of our observations into a consistent photometric system, we attempted to make sure that there were overlap observations between runs, and we developed a pair of programs called ``apphot'' and ``apcomp'' to compare observations. ``Apphot'' converts a galaxy image with its photometric calibration into a table of circular aperture photometry. This only depends on plate scale (which is well known) and therefore permits comparison of different images regardless of their angular orientation. ``Apcomp'' then takes the aperture photometry from two observations and fits the two profiles to one another using a photometry scale offset and a relative sky level. These two programs can give us accurate offsets between the photometry of two images, good at the 0.005 magnitude level. We learned, however, that good seeing is much more common than photometric weather, and we realized that many of our ``photometric'' observations were not reliable at the 0.01 magnitude level. As the survey progressed and the number of overlaps increased, we also realized that although we only need 0.05 magnitude photometry of \ifmmode\overline{m}_I\else$\overline{m}_I$\fi, \ifmmode\overline{M}_I\else$\overline{M}_I$\fi\ is sensitive enough to \ifmmode(V{-}I)_0\else$(V{-}I)_0$\fi\ color that we needed better photometry. The existing photoelectric (PE) photometry, although probably very good in the mean, is neither extensive enough nor accurate enough to serve to calibrate the survey. We also became aware that there are many peculiar CCD and shutter effects which make good photometry difficult. For example, we have found photometry with Tektronix (SITe) CCDs particularly challenging for reasons we do not fully understand. Because of their high quantum efficiency and low noise they have been the detectors of choice, but run to run comparisons with apphot and apcomp show consistent zero point offsets at the 0.05 magnitude level. While not a serious problem for \ifmmode\overline{m}_I\else$\overline{m}_I$\fi, we had to do much better in measuring \ifmmode(V{-}I)_0\else$(V{-}I)_0$\fi. Accordingly, we undertook an auxiliary survey in 1995 of a substantial fraction of our SBF survey from the McGraw-Hill 1.3-m telescope at the MDM Observatory. We shared the time with another program and used only nights which were photometric, as judged by the observer at the time and as revealed later from the quality of the photometric standard observations. We did not use Tek CCDs but primarily used the front-illuminated, Loral 2048$^2$ CCD Wilbur (Metzger et al. 1993), we used filters which match $V$ and $I_{KC}$ as closely as possible, and the large field of view permitted us to make very good estimates of sky levels. Over 5 runs this comprised about 600 observations in $V$ and $I$. We made certain to have a generous overlap between these observations and all our other observing runs, reaching well south to tie to the CTIO and LCO data. We then performed a grand intercomparison of all the photometric data in order to determine photometric offsets from run to run. Using apphot and apcomp, we determined offsets between observations, and we built up a large table of comparison pairs. In addition, photoelectric (PE) photometry from deVaucouleurs and Longo (1988), Poulain and Nieto (1994), and Buta and Williams (1995) served as additional sources of comparison, and we computed differences between PE and our photometry for every galaxy in common. We have found that \ifmmode(V{-}I)_0\else$(V{-}I)_0$\fi\ colors often show somewhat better agreement than the individual $V$ and $I$ measurements, presumably because thin clouds are reasonably gray, so we also compared colors directly in addition to photometric zero points. The results are illustrated in Figure 1. In each of three quantities $V$, $I$, and \ifmmode(V{-}I)_0\else$(V{-}I)_0$\fi, we fitted for zero point offsets for each run (photoelectric sources were considered to be a ``run''), minimizing the pairwise differences. We set the overall zero point by insisting that the median run offset be zero. Upon completion, we found that the rms of the zero point offsets to be 0.029 mag, and the rms scatter of individual comparisons between CCD data to be 0.030 mag in $V$, 0.026 mag in $I$, and 0.024 mag in \ifmmode(V{-}I)_0\else$(V{-}I)_0$\fi. The scatter was bigger for CCD-PE zero point comparisons, 0.047 mag in both $V$ and $I$. The ``zero point offsets'' for the photoelectric photometry were 0.003 mag in $V$, 0.017 mag in $I$, and 0.004 mag in \ifmmode(V{-}I)_0\else$(V{-}I)_0$\fi, which we take to be close enough to zero that we did not choose to modify our overall median zero point to force them to zero. Finally, we chose zero point corrections for $V$ and $I$ for each run according to these offsets. The difference of the corrections was set to the \ifmmode(V{-}I)_0\else$(V{-}I)_0$\fi\ offset from the comparison, and the sum of the corrections was the sum of the $V$ and $I$ offsets. We therefore believe that (a) our photometry is very close to Landolt and photoelectric in zero point, (b) the error in the $V$ or $I$ photometry for a given observation is 0.02 mag, and (c) the error in a given \ifmmode(V{-}I)_0\else$(V{-}I)_0$\fi\ color measurement is 0.017 mag (where we have divided by $\sqrt2$ to get the error for single measurements). We also add in quadrature 0.25 of the zero point offset which was applied. The offsets $\Delta V$ and $\Delta I$ for each run are listed in Table 1. The reductions of \ifmmode\overline{m}_I\else$\overline{m}_I$\fi\ are described elsewhere (e.g., Jacoby et al. 1992). Briefly, we fit a galaxy model to the summed, cleaned, sky-subtracted galaxy image and subtract it. If there is dust present (all too common in ellipticals and S0s as well as the large bulge spiral galaxies we observe), we mask it out as well. Experiments with masking different portions of M31 and M81 (where we used $B$ band observations to show us clearly the location of the dust) indicate that reasonable care in excising dust will produce a reliable \ifmmode\overline{m}_I\else$\overline{m}_I$\fi, both because the extinction is less in the $I$ band and also because the dust tends to cause structure at relatively large scales which are avoided by our fit to the Fourier power spectrum. We run DoPhot (Schechter et al. 1993) on the resulting image to find stars, globular clusters, and background galaxies; fit a luminosity function to the results; and derive a mask of objects brighter than a completeness limit and an estimate of residual variance from sources fainter than the limit. Applying the mask to the model-subtracted image, we calculate the variance from the fluctuations in a number of different regions. Finally, this variance is converted to a value for \ifmmode\overline{m}_I\else$\overline{m}_I$\fi\ by dividing by the mean galaxy flux and subtraction of the residual variance estimate from unexcised point sources. Generally speaking, the various estimates of \ifmmode\overline{m}_I\else$\overline{m}_I$\fi\ are quite consistent from region to region, and a weighted average and error estimate are tabulated for each observation. If the observation was photometric, we also record the \ifmmode(V{-}I)_0\else$(V{-}I)_0$\fi\ color found in the same region in which we measure \ifmmode\overline{m}_I\else$\overline{m}_I$\fi. There are many galaxies for which \ifmmode(V{-}I)_0\else$(V{-}I)_0$\fi\ and \ifmmode\overline{m}_I\else$\overline{m}_I$\fi\ have been measured more than once, and intercomparison of the different observations can be used to evaluate whether our error estimates are reasonable. If we consider all pairs and divide their difference by the expected error, the distribution should be a Gaussian of unity variance. Figure 2 illustrates these distributions for \ifmmode\overline{m}_I\else$\overline{m}_I$\fi\ and \ifmmode(V{-}I)_0\else$(V{-}I)_0$\fi. Evidently, the error estimates are usually quite good, with discordant observations occurring rarely. In most cases of discordances, it is clear which of the observations is trustworthy, and we simply remove the other observation from further consideration. These excised observations occur 1.5\% of the time for \ifmmode\overline{m}_I\else$\overline{m}_I$\fi\ and 0.3\% of the time for \ifmmode(V{-}I)_0\else$(V{-}I)_0$\fi, and are an indication of how frequently bad observations occur. After observations are averaged together, they are subjected to some final corrections. The mean \ifmmode(V{-}I)\else$(V{-}I)$\fi\ color of the fluctuations is the mean of a galaxy's color \ifmmode(V{-}I)\else$(V{-}I)$\fi\ and the ``fluctuation color'' \ifmmode\overline{m}_V\else$\overline{m}_V$\fi--\ifmmode\overline{m}_I\else$\overline{m}_I$\fi, or $\ifmmode(V{-}I)\else$(V{-}I)$\fi\approx 1.85$ (since the rms fluctuation is the square root of the flux from the galaxy and the flux from \ifmmode\overline{m}_I\else$\overline{m}_I$\fi). The value of \ifmmode\overline{m}_I\else$\overline{m}_I$\fi\ is corrected according to this mean color and the color term for the run's photometry. The values of \ifmmode\overline{m}_I\else$\overline{m}_I$\fi\ and \ifmmode(V{-}I)\else$(V{-}I)$\fi\ are corrected for galactic extinction according to \begin{equation} A_V:A_{I_{\rm KC}}:E(B-V) = 3.04:1.88:1.00, \end{equation} where $E(B-V)$ comes from Burstein \& Heiles (1984), who give $A_B = 4.0\, E(B-V)$, the relative extinction ratio $A_{I_{\rm KC}} / A_V = 0.62$ is taken from Cohen et al. (1981) for a star halfway between an A0 and an M star, and $A_V/E(B-V)$ is an adjustment of a value of 3.1 for A0 stars common in the literature ({\it e.g.,} Cardelli, et al. 1989) to a value of 3.04 more appropriate for early-type galaxies, following the ratios given in Cohen et al. The final modification is the application of K-corrections which brighten magnitudes in $V$ and $I$ by 1.9 and $1.0 \times z$ respectively (Schneider 1996), and brighten fluctuation magnitudes in $I$ by $7.0 \times z$ (Worthey 1996). Note that the very red color of SBF causes flux to be shifted rapidly out of the $I$ band with redshift, but the \ifmmode\overline{m}_I\else$\overline{m}_I$\fi\ K-correction amounts to only 0.05 magnitude at a typical distance of 2000~km/s. \section{Calibrating \ifmmode\overline{M}_I\else$\overline{M}_I$\fi} The next step we take in trying to establish how \ifmmode\overline{M}_I\else$\overline{M}_I$\fi\ varies according to stellar population is to look at how \ifmmode\overline{m}_I\else$\overline{m}_I$\fi\ varies from galaxy to galaxy within groups, where the distance to the galaxies is essentially constant. We originally chose to observe SBF in the $I$ band because stellar population models indicate that \ifmmode\overline{M}_I\else$\overline{M}_I$\fi\ is relatively constant from population to population, and that the effects of age, metallicity, and IMF are almost degenerate --- in other words, \ifmmode\overline{M}_I\else$\overline{M}_I$\fi\ is nearly a one parameter family. Guided by theoretical models we seek to establish whether three statements are a fair description of our data: \par\indent \hangindent2\parindent \textindent{(\S3.1)} \ifmmode\overline{M}_I\else$\overline{M}_I$\fi\ is a one-parameter family, with a universal dependence on \ifmmode(V{-}I)_0\else$(V{-}I)_0$\fi\break (i.e., \ifmmode\overline{M}_I\else$\overline{M}_I$\fi\ is a function of \ifmmode(V{-}I)_0\else$(V{-}I)_0$\fi\ with small residual scatter). \par\indent \hangindent2\parindent \textindent{(\S3.2)} The zero point of the \ifmmode\overline{M}_I\else$\overline{M}_I$\fi--\ifmmode(V{-}I)_0\else$(V{-}I)_0$\fi\ relation is universal. \par\indent \hangindent2\parindent \textindent{(\S3.3)} The \ifmmode\overline{M}_I\else$\overline{M}_I$\fi--\ifmmode(V{-}I)_0\else$(V{-}I)_0$\fi\ relation is consistent with theoretical models of stellar populations. To this end we chose approximately 40 nearby groups where we currently have (or will have) observed more than one galaxy. The groups are defined by position on the sky and a redshift range and in most cases correspond to one of the groups described by Faber et al. (1989). Table 3 lists our groups. Note that we are not trying to include all groups, nor do we have to be complete in including all galaxies which are members. We are simply trying to create samples of galaxies for which we are reasonably confident that all galaxies are at the same distance. \subsection{Universality of the \ifmmode\overline{m}_I\else$\overline{m}_I$\fi\ dependence on \ifmmode(V{-}I)_0\else$(V{-}I)_0$\fi} Figure 3 illustrates the \ifmmode\overline{m}_I\else$\overline{m}_I$\fi--\ifmmode(V{-}I)_0\else$(V{-}I)_0$\fi\ relationship in six groups where we have measured SBF in a number of galaxies: NGC~1023, Leo, Ursa Major, Coma~I\&II, Virgo, and Fornax. The lines are drawn with slope 4.5 and zero point according to the fit to the data described below. We see that galaxies which meet the group criteria of position on the sky and redshift are consistent with the same \ifmmode\overline{m}_I\else$\overline{m}_I$\fi--\ifmmode(V{-}I)_0\else$(V{-}I)_0$\fi\ relationship, where the scatter reflects both the measurement error and the group depth inferred from spread across the sky. In Virgo we find NGC~4600 much brighter than the rest of the galaxies, NGC~4365 significantly fainter, and NGC~4660 (the point at $\ifmmode(V{-}I)_0\else$(V{-}I)_0$\fi=1.21$ and $\ifmmode\overline{m}_I\else$\overline{m}_I$\fi = 28.9$) also with an unusually bright \ifmmode\overline{m}_I\else$\overline{m}_I$\fi\ for its color. These three galaxies, marked as smaller, square symbols, are discussed below. Note that ellipticals and S0 galaxies are intermixed with spirals (NGC~3368 in Leo, NGC~4548 in Virgo, NGC~891 in the NGC~1023 group, and NGC~4565 and NGC~4725 in the Coma~I\&II group). The two galaxies in Fornax marked as ``spiral'' (NGC~1373 and NGC~1386) might better be classified as S0 on our deep CCD images. For this admittedly small sample we see no offset between SBF measurements in spiral bulges and early-type galaxies. We regard this as confirmation of our assumption that SBF measurements are equally valid in spiral bulges as in early-type galaxies. In order to test the hypothesis that \ifmmode\overline{M}_I\else$\overline{M}_I$\fi\ has a universal dependence on \ifmmode(V{-}I)_0\else$(V{-}I)_0$\fi\ in a more systematic way than fitting individual groups, we simultaneously fit all our galaxies which match the group criteria with \begin{equation} \overline m_I = \langle\overline m_I^0\rangle_j + \beta \; [\ifmmode(V{-}I)_0\else$(V{-}I)_0$\fi-1.15], \end{equation} where we fit for values of $\langle\overline m_I^0\rangle_j$ for each of j=1,N groups and a single value for $\beta$. The quantity $\langle\overline m_I^0\rangle_j$ is the group mean value for \ifmmode\overline{m}_I\else$\overline{m}_I$\fi\ at a fiducial galaxy color of $\ifmmode(V{-}I)_0\else$(V{-}I)_0$\fi=1.15$. The measurements of \ifmmode(V{-}I)_0\else$(V{-}I)_0$\fi\ and \ifmmode\overline{m}_I\else$\overline{m}_I$\fi\ carry errors which the pair-wise comparisons and the averaging procedure of section 2 indicate are accurate. We also anticipate that there will be an irreducable ``cosmic'' scatter in \ifmmode\overline{M}_I\else$\overline{M}_I$\fi. Accordingly, in fitting \ifmmode\overline{m}_I\else$\overline{m}_I$\fi\ as a function of \ifmmode(V{-}I)_0\else$(V{-}I)_0$\fi, we include an error allowance for this cosmic scatter which is nominally 0.05 magnitudes (i.e. for this fit the error in \ifmmode\overline{m}_I\else$\overline{m}_I$\fi\ is enhanced by 0.05 magnitude in quadrature). In addition, we will also see scatter because the galaxies within groups are not truly at the same distance. We therefore calculate the rms angular position of the galaxies making up each group, and divide this radius by $\sqrt2$ as an estimate of the rms group depth. Converting this to a magnitude, we add it in quadrature to the error in \ifmmode\overline{m}_I\else$\overline{m}_I$\fi. We then perform a linear fit of $N+1$ parameters which allows for errors in both the ordinate and abcissa, according to the ``least distance'' method used by Tonry and Davis (1981). (This also appears in a slightly different guise in the second edition of {\it Numerical Recipes} by Press et al. 1992) We remove the three Virgo galaxies which we believe are at significantly different distances from the rest of the group (NGC~4365, NGC~4660, and NGC~4600), mindful that what is considered to be part of Virgo and what is not is somewhat arbitrary. We also choose to exclude NGC~205 and NGC~5253 from the fit because recent starbursts make them extremely blue --- we do not believe our modeling extends to such young populations. With 149 galaxies we have 117 degrees of freedom, and we find that $\chi^2 = 129$, $\chi^2/N = 1.10$, and the slope of the \ifmmode\overline{M}_I\else$\overline{M}_I$\fi--\ifmmode(V{-}I)_0\else$(V{-}I)_0$\fi\ relation is $4.5 \pm 0.25$. The galaxies contributing to the fit span a color range of $1.0<\ifmmode(V{-}I)_0\else$(V{-}I)_0$\fi<1.3$. Because Virgo still contributes five of the seven most discrepant points (the other two are in Cetus), the rms depth used for Virgo ($2.35^\circ \rightarrow 0.08$~mag) may be too small, making $\chi^2/N$ slightly bigger than one. If we replace the 3 Virgo galaxies we omitted earlier, we find that $\chi^2/N$ rises to 1.75 for 120 degrees of freedom and the slope changes to $4.7\pm0.25$, showing that even though these galaxies are significantly outside of Virgo, the slope is robust. When we experiment with adding and removing different groups we find that the slope changes slightly, but is always consistent with the error above. These values for $\chi^2$ include an allowance for cosmic scatter of 0.05 magnitude and the nominal, rms group depth. These two, ill-constrained sources of error can play off against each other: if we double the group depth error allowance, we get $\chi^2/N = 1.0$ for zero cosmic scatter; if we increase the cosmic scatter to 0.10 magnitude, we need to decrease the group depth to zero in order to make $\chi^2/N = 1.0$. Therefore, even though we cannot unambiguously determine how much cosmic scatter there is in the \ifmmode\overline{M}_I\else$\overline{M}_I$\fi--\ifmmode(V{-}I)_0\else$(V{-}I)_0$\fi\ relation, it appears to be $\sim$0.05 mag. The referee pointed out that even if we make no allowance for group depth, the cosmic scatter of 0.10 mag makes SBF the most precise tertiary distance estimator by far, and wanted to know how sensitive this is to our estimates of observational error. There is not much latitude for the cosmic scatter to be larger than 0.10 mag. The distribution of measurement error in \ifmmode\overline{m}_I\else$\overline{m}_I$\fi\ and \ifmmode(V{-}I)_0\else$(V{-}I)_0$\fi\ (which also enter $\chi^2$) starts at 0.06 mag, and has quartiles at 0.11, 0.16, and 0.20 mag. If we wanted to increase the cosmic scatter by $\sqrt2$ to 0.14 mag, we would have to have overestimated the observational errors by 0.10 mag in quadrature, and apart from the fact that a quarter of the measurements would then have imaginary errors, our pairwise comparison of multiple observations from the previous section would not allow such a gross reduction in observational error. Figure 4 illustrates how \ifmmode\overline{m}_I\else$\overline{m}_I$\fi\ depends on \ifmmode(V{-}I)_0\else$(V{-}I)_0$\fi\ when all the group data have been slid together by subtraction of the group mean at $\ifmmode(V{-}I)_0\else$(V{-}I)_0$\fi=1.15$. Note again that spiral galaxies, in this case four galaxies with both Cepheid and SBF distances, show no offset relative to the other early-type galaxies making up the groups in which they appear, other than the usual trend with \ifmmode(V{-}I)_0\else$(V{-}I)_0$\fi. The overall rms scatter, 0.18 mag, arising from all the effects discussed above, is a testament to the quality of SBF as a distance estimator. The Local Group galaxies NGC~205, NGC~147, and NGC~185 have also been plotted in Figure 4 (although they were not used in the fit), under the assumption that they are at the same distance as M31 and M32. This may or may not be a valid assumption for NGC~147 and NGC~185, but they agree reasonably well with the mean relation. In contrast to these two galaxies, which are blue because of extremely low metallicity, NGC~205 has undergone a recent burst of star formation and has a strong A star spectral signature. Because our models do {\it not} extend to such young populations, the systematic deviation from the mean relation is not unexpected. The inset in Figure 4 extends the color range to show that this deviation continues for two other galaxies where there has been recent star formation: NGC~5253 and IC~4182. NGC~5253 is 0.5 mag fainter than one would expect using a naive extrapolation of the relation to its color of $\ifmmode(V{-}I)_0\else$(V{-}I)_0$\fi = 0.84$, and IC~4182 has an SBF magnitude which is 0.75 mag fainter than one would judge from its Cepheid distance and its color of $\ifmmode(V{-}I)_0\else$(V{-}I)_0$\fi = 0.71$. Qualitatively this makes sense because the very young stars change the overall color of the galaxy quite a bit but are not very luminous in the $I$ band compared to the stars at the top of the RGB which are the main contributors to the SBF \ifmmode\overline{m}_I\else$\overline{m}_I$\fi. It may be that these very young populations can be understood well enough that one can safely predict the SBF absolute magnitude from the mean color, but this is beyond the scope of this paper. Tammann (1992) expressed concern that there are residual stellar population effects in SBF even after the correction for \ifmmode(V{-}I)_0\else$(V{-}I)_0$\fi\ color. However, his critique was based on an early attempt to correlate \ifmmode\overline{m}_I\else$\overline{m}_I$\fi\ as a function of \ifmmode(V{-}I)_0\else$(V{-}I)_0$\fi\ (Tonry et al. 1990). Unfortunately, that work had the wrong sign for the slope (appropriate for the $JHK$ bandpasses but not $I$), because it was based on the Revised Yale Isochrones (Green et al. 1987), which did not properly model the line blanketing in metal rich, high luminosity stars. The effect noted by Tammann was a residual correlation of the corrected \ifmmode\overline{m}_I^0\else$\overline{m}_I^0$\fi\ with the Mg$_2$ index among galaxies within a cluster. Figure 5 shows these trends do not exist for the present data and the new \ifmmode\overline{m}_I\else$\overline{m}_I$\fi--\ifmmode(V{-}I)_0\else$(V{-}I)_0$\fi\ relation: in both Fornax and Virgo there is no residual correlation with either Mg$_2$ or galaxy magnitude. We conclude that a one-parameter, linear relation between \ifmmode\overline{M}_I\else$\overline{M}_I$\fi\ and \ifmmode(V{-}I)_0\else$(V{-}I)_0$\fi\ suffices to describe our data for $1.0 < \ifmmode(V{-}I)_0\else$(V{-}I)_0$\fi < 1.3$; the slope of the \ifmmode\overline{M}_I\else$\overline{M}_I$\fi--\ifmmode(V{-}I)_0\else$(V{-}I)_0$\fi\ relation is universally $4.5 \pm 0.25$, and we are indeed detecting cosmic scatter in \ifmmode\overline{M}_I\else$\overline{M}_I$\fi\ of order 0.05 mag. Very few galaxies fail to follow the relation, and for every such galaxy at least one of the following statements is true: (1) the measurement of \ifmmode\overline{m}_I\else$\overline{m}_I$\fi\ or \ifmmode(V{-}I)_0\else$(V{-}I)_0$\fi\ is doubtful; (2) the galaxy may not be a member of the group we assigned it to; (3) the stellar population is bluer than $\ifmmode(V{-}I)_0\else$(V{-}I)_0$\fi=1.05$ due to recent star formation. Note that this slope is steeper than the value of 3 tendered by Tonry (1991) and used by Ciardullo et al. (1993) who suggested that it might be as steep as 4. Basically, the reason for this is that the older data were noiser and were fitted only to errors in the ordinate, whereas in fact the errors in \ifmmode(V{-}I)_0\else$(V{-}I)_0$\fi\ are quite significant, particularly for the better measured \ifmmode\overline{m}_I\else$\overline{m}_I$\fi, which count heavily in any weighted fit. \subsection{Universality of the \ifmmode\overline{M}_I\else$\overline{M}_I$\fi\ zero point} We have effectively tested the hypothesis that the zero point of SBF is universal within groups, but in order to extend the test from group to group we need independent distance estimates. Since the groups are all nearby, the group's redshift is not an accurate distance estimate --- there are likely to be substantial non-Hubble velocities included in the group's recession velocity. We therefore turn to other distance estimators: Cepheids, planetary nebula luminosity function (PNLF), Tully-Fisher (TF), \ifmmode D_n{-}\sigma\else$ D_n{-}\sigma$\fi, Type II supernovae (SNII) and Type I supernovae (SNIa). Some of these estimators have zero points in terms of Mpc (such as Cepheids and SNII), others have zero points in terms of km/s based on the Hubble flow (such as \ifmmode D_n{-}\sigma\else$ D_n{-}\sigma$\fi), and a few have both (such as TF). For our initial discussion we seek only to establish whether the relative distances agree with SBF; for now we do not care about the zero point, though it will soon be addressed. Figures 6 and 7 show the comparison between the values of the SBF parameters \ifmmode\langle\overline{m}_I^0\rangle\else$\langle\overline{m}_I^0\rangle$\fi\ derived previously for each of our groups and the distances to the groups according to these 6 methods. The results of fitting lines of unity slope (allowing for errors in both coordinates) to the data in each panel are given in Table 4. We use the published error estimates for all of these other indicators so $\chi^2/N$ should be viewed with some caution: outliers and non-Gaussian errors or over-optimistic error estimates can inflate $\chi^2/N$ even though the mean offset is still valuable. Since each comparison is very important, we briefly discuss them individually. \subsubsection{Cepheids} There is now a growing number of Cepheid distances with which we compare, but we are faced with the complication that Cepheids occur in young stellar populations, while SBF is best measured where such populations are not present. There are five galaxies which have both Cepheid and SBF distances: NGC~224 (Freedman \& Madore 1990), NGC~3031 (Freedman et al 1994), NGC~3368 (Tanvir et al. 1995), NGC~5253 (Saha et al. 1995), and NGC~7331 (Hughes 1996). NGC~5253 is especially problematic for SBF, because its recent starburst has produced a much younger and bluer stellar population than we have calibrated. We can, of course, also compare distances according to group membership. There are 7 groups where this is currently possible: Local Group, M81, CenA, NGC~1023 (NGC~925 from Silbermann et al. 1996), NGC~3379 (also including NGC~3351 from Graham et al. 1996), NGC~7331, and Virgo (including NGC~4321 from Ferrarese et al. 1996, NGC~4536 from Saha et al. 1996a, and NGC~4496A from Saha et al. 1996b; we exclude NGC~4639 from Sandage et al. 1996 because we are also excluding NGC~4365 and the W cloud from the SBF mean). In the former case we find that fitting a line to \ifmmode\langle\overline{m}_I^0\rangle\else$\langle\overline{m}_I^0\rangle$\fi\ as a function of \ifmmode(m{-}M)\else$(m{-}M)$\fi\ yields a mean offset of $-1.75\pm0.05$ mag with $\chi^2/N$ of 3.4 for 4 degrees of freedom, and $-1.82\pm0.06$ mag with $\chi^2/N$ of 0.3 for 3 degrees of freedom when NGC~5253 is excluded. In the latter case we get a mean offset of $-1.74\pm0.05$ mag with $\chi^2/N$ of 0.6 for 6 degrees of freedom. When NGC~5253 is excluded, the rms scatter is remarkably small, only 0.12 magnitudes for the galaxy comparison and 0.16 magnitudes for the group comparison. \subsubsection{PNLF} Ciardullo et al. (1993) reported virtually perfect agreement between SBF and PNLF, but recent publications (Jacoby et al. 1996) have raised some discrepancies. Examination of Figure 7 reveals that our fit has two outliers: Coma~I (e.g. NGC~4278) and Coma~II (e.g. NGC~4494). Because we do not know how to resolve this issue at present, Table 4 gives the result for $\ifmmode\langle\overline{m}_I^0\rangle\else$\langle\overline{m}_I^0\rangle$\fi - \ifmmode(m{-}M)\else$(m{-}M)$\fi_{\hbox{PNLF}}$ for the entire sample and when these two outliers are removed. Since PNLF is fundamentally calibrated on Cepheids, this is not independent of the previous number, but it does confirm that PNLF and SBF are measuring the same relative distances. \subsubsection{SNII} The expanding photospheres method (EPM) described most recently by Eastman et al. (1996) offers distance estimates which are largely independent of the Cepheid distance scale. There is only one galaxy with both an EPM and an SBF distance (NGC~7331), but there have also been two SNII in Dorado (NGC~1559 and NGC~2082), two in Virgo (NGC~4321 and NGC~4579), and one in the NGC~1023 group (NGC~1058). The agreement between EPM and SBF (Fig. 6) is good. The farthest outlier is NGC~7331, for which SBF and Cepheid distances are discordant with the SNII distance. Table 4 lists separately the zero point, scatter, and $\chi^2/N$ when NGC~7331 is included and excluded. \subsubsection{TF (Mpc calibration)} B. Tully (1996) was kind enough to provide us with TF distances to the SBF groups in advance of publication. The fit between TF and SBF gives $\ifmmode\langle\overline{m}_I^0\rangle\else$\langle\overline{m}_I^0\rangle$\fi - \ifmmode(m{-}M)\else$(m{-}M)$\fi = -1.69\pm0.03$ mag. This is again not independent of the Cepheid number, since the TF zero point comes from the same Cepheid distances. Figure 7 demonstrates that the agreement is generally good, despite the high $\chi^2/N$ which comes from a few non-Gaussian outliers. We cannot tell whether these outliers reflect non-Gaussian errors in the methods or simply the difficulties of choosing spirals and early type galaxies in the same groups. \subsubsection{TF (km/s calibration)} We applied the SBF group criteria to the ``Mark II'' catalog of galaxy distances distributed by D. Burstein. We selected all galaxies with ``good'' TF distances (mostly from Aaronson et al. 1982) and computed an average distance to the groups, applying the usual Malmquist bias correction according to the precepts of Lynden-Bell et al. (1988) and the error estimates from Burstein. Because these distances have a zero point based on the distant Hubble flow, we derive an average offset of $\ifmmode\langle\overline{m}_I^0\rangle\else$\langle\overline{m}_I^0\rangle$\fi\ - 5\log d = 13.55\pm0.08$ mag. \subsubsection{\ifmmode D_n{-}\sigma\else$ D_n{-}\sigma$\fi} Most of the SBF groups are the same as those defined by Faber et al. (1989). We compare their Malmquist bias corrected distances to these groups (which are based on a zero point from the distant Hubble flow) with SBF and find the same result as Jacoby et al. (1992): the distribution of errors has a larger tail than Gaussian, but the error estimates accurately describe the central core of the distribution. $\chi^2/N$ is distinctly larger than 1, but the difference histogram in Figure 7 reveals that this is because of the tails of the distribution. The fit between \ifmmode D_n{-}\sigma\else$ D_n{-}\sigma$\fi\ and SBF gives $\ifmmode\langle\overline{m}_I^0\rangle\else$\langle\overline{m}_I^0\rangle$\fi\ - 5\log d = 13.64\pm0.05$ mag. \subsubsection{SNIa} Extraordinary claims have been made recently about the quality of SNIa as distance estimators. Some authors (e.g. Sandage and Tammann 1993) claim that suitably selected (``Branch normal'') SNIa are standard candles with a dispersion as little as 0.2 mag. Others (e.g. Phillips 1993) believe that they see a correlation between SNIa luminosity and their rate of decline, parametrized by the amount of dimming 15 days after maximum, $\Delta m_{15}$. Still others (e.g. Riess et al. 1995) agree with Phillips (1993) but believe that they can categorize SNIa better by using more information about the light curve shape than just this rate of decline. Finally, there is the ``nebular SNIa method'' of Ruiz-Lapuente (1996) which tries to determine the mass of the exploding white dwarf by consideration of the emission lines from the expanding ejecta. We therefore choose to compare SBF distances with SNIa under two assumptions: that SNIa are standard candles, and that $m_{max} - \alpha\Delta m_{15}$ is a better indicator of distance. In both cases we restrict our fits to $0.8 < \Delta m_{15} < 1.5$ as suggested by Hamuy et al. (1995) and use a distance error of 0.225 mag for each SNIa. SNIa have been carefully tied to a zero point according to the distant Hubble flow (one of the main advantages of SNIa) by Hamuy et al. (1995), under both assumptions. There have also been vigorous attempts to tie the SNIa to the Cepheid distance scale which we have chosen not to use because of the circularity with our direct comparison between SBF and Cepheids. The results are both encouraging and discouraging. We find that there is indeed a good correlation between SNIa distance and SBF, with average values of $\ifmmode\langle\overline{m}_I^0\rangle\else$\langle\overline{m}_I^0\rangle$\fi\ - 5\log d = 13.92\pm0.08$ mag and $14.01\pm0.08$ mag for the group comparison under the two assumptions. As illustrated in Figure 6, $m_{max} - \alpha\Delta m_{15}$ does correlate better with distance than $m_{max}$, but as long as ``fast declining'' SNIa are left out there is scant difference between the zero point according to the two methods. The panels of Figure 6 showing SBF and SNIa hint at a systematic change between the nearest three and the farthest three groups, in the sense that there appears to be a change in zero point by about 0.7 mag. One might worry that this is evidence that SBF is ``bottoming out'', but there is no hint of this in the comparisons with TF and \ifmmode D_n{-}\sigma\else$ D_n{-}\sigma$\fi\ in Figure 7 which extend to much fainter \ifmmode\overline{m}_I\else$\overline{m}_I$\fi. One might also worry about whether there are systematic differences in SNIa in spirals and ellipticals, and biases from the lack of nearby ellipticals or S0 galaxies. However, it is probably premature to examine these points in too much detail. For example, the point at $\ifmmode\langle\overline{m}_I^0\rangle\else$\langle\overline{m}_I^0\rangle$\fi \approx 28$ uses the SBF distance to Leo~I, but the SNIa occurred in NGC~3627 which lies $8^\circ$ away from the Leo~I group. This is a fundamental difficulty in the SBF--SNIa comparison, which will improve as SBF extends to greater distances and more nearby SNIa are observed. There are seven galaxies bearing SNIa where SBF distances have been measured: NGC~5253 (SN~1972E), NGC~5128 (SN~1986G), NGC~4526 (SN~1994D), NCG~2962 (SN~1995D), NGC~1380 (SN~1992A), NGC~4374 (SN~1991bg), NGC~1316 (SN~1980N). Inasmuch as two of these are slow decliners (SN~1986G, SN~1991bg), we fit the remaining five using the SBF distance to the galaxy instead of the group. We derive $\ifmmode\langle\overline{m}_I^0\rangle\else$\langle\overline{m}_I^0\rangle$\fi\ - 5\log d = 13.86\pm0.12$ mag and $14.01\pm0.12$ mag for the two methods. We regard the SBF distance to NGC~5253 as uncertain because we have not calibrated \ifmmode\overline{M}_I\else$\overline{M}_I$\fi\ for such a young stellar population. We thus also recompare SBF and SNIa with NGC~5253 removed from consideration. $\chi^2/N$ becomes dramatically smaller in both cases and $\ifmmode\langle\overline{m}_I^0\rangle\else$\langle\overline{m}_I^0\rangle$\fi\ - 5\log d$ become smaller by about 0.2 mag to $13.64\pm0.13$ mag and $13.87\pm0.13$ mag. \subsubsection{Zero point summary} These comparisons demonstrate that the second hypothesis is correct: the zero point of the \ifmmode\overline{M}_I\else$\overline{M}_I$\fi--\ifmmode(V{-}I)_0\else$(V{-}I)_0$\fi\ relationship is universal. We use the SBF--Cepheid fit to derive a final, empirical relationship between \ifmmode\overline{m}_I\else$\overline{m}_I$\fi\ and \ifmmode(V{-}I)_0\else$(V{-}I)_0$\fi: \begin{equation} \overline M_I = (-1.74\pm0.07)\, +\, (4.5\pm0.25)\, [\ifmmode(V{-}I)_0\else$(V{-}I)_0$\fi - 1.15]. \end{equation} This zero point differs from that of Tonry (1991) by about 0.35 magnitude. The reason is simply that the 1991 zero point was based entirely on M31 and M32, and the observational error in both \ifmmode\overline{m}_I\else$\overline{m}_I$\fi\ and \ifmmode(V{-}I)_0\else$(V{-}I)_0$\fi\ worked in the same direction, as did the photometric zero point errors (cf. Table 1 for K0990). The SBF distances which have been published therefore increase by about 15 percent (for example Fornax moves from 15~Mpc to 17~Mpc), except for Virgo, where the earlier result included NGC~4365 which we now exclude in calculating the average distance to the core of the cluster. This new calibration is based on 10 Cepheid distances in 7 groups and 44 SBF distances. As seen in Figure 6 and Table 4, these are highly consistent with one another with a scatter of about 0.15 mag. Along with the extensive photometric recalibration, this zero point should be accurate to $\pm0.07$ mag. This error estimate makes an allowance of 0.05 mag for the uncertainty in the Cepheid zero point in addition to the statistical error of 0.05 mag, and the comparisons with theory and SNII give us confidence that this truly is correct. \subsection{Comparison with theory} Finally we test our third hypothesis by comparing our \ifmmode\overline{M}_I\else$\overline{M}_I$\fi--\ifmmode(V{-}I)_0\else$(V{-}I)_0$\fi\ relationship with theoretical models of stellar populations. Figure 8 shows the model predictions of Worthey (1993a,b) along with the empirical line. When the theoretical models are fitted with the empirically determined slope of 4.5, they yield a theoretical zero point of $-1.81$ mag with an rms scatter of $0.11$ mag for the SBF relation. We enter this value in Table 4, with the scatter offered as an ``error estimate'', but it must be remembered that this is fundamentally different from the other entries in the table. There is good agreement here, although the theoretical result for \ifmmode\overline{M}_I\else$\overline{M}_I$\fi\ may be slightly brighter (0.07 mag) or slightly redder (0.015 mag) than the empirical result. Given the difficulties that the theoretical models have in simultaneously fitting the color and Mg$_2$ indices of real galaxies, we regard this agreement as excellent confirmation of the empirical calibration. \section{The Hubble Constant} The scope of this paper does not extend to comparing SBF distances with velocity; this will be the subject of the next paper in the series. However, the comparison with other distance estimators does provide us with a measurement of the Hubble constant. The comparison with other estimators whose zero point is defined in terms of Mpc tells us the absolute magnitude of SBF. At our fiducial color of $\ifmmode(V{-}I)_0\else$(V{-}I)_0$\fi=1.15$, we find that Cepheids give us an absolute magnitude $\ifmmode\overline{M}_I\else$\overline{M}_I$\fi = -1.74\pm0.05$. We prefer the group-based Cepheid comparison because of the very few SBF measurements possible in spirals which have Cepheids. The other Mpc-based distance estimators are all consistent with this zero point, as we would hope since they are calibrated with the same Cepheid data. The results from theoretical models of stellar populations and SNII are also consistent with this zero point, and provide independent confirmation of the validity of the Cepheid distance scale. The comparison of SBF with estimators whose zero point is based on the large scale Hubble flow is less consistent. The estimators based on galaxy properties, TF and \ifmmode D_n{-}\sigma\else$ D_n{-}\sigma$\fi, are consistent with one another and consistent with SBF in terms of relative distances. They give a zero point for SBF at the fiducial color of $\ifmmode(V{-}I)_0\else$(V{-}I)_0$\fi=1.15$ of $\ifmmode\langle\overline{m}_I^0\rangle\else$\langle\overline{m}_I^0\rangle$\fi = 5\log d\hbox{(km/s)} + 13.59\pm0.07$, where the error comes from the rms divided by $\sqrt{N-1}$. Supernovae and SBF are more interesting. The group membership of the Cepheid galaxies was not difficult since they were specifically chosen to be group members. In contrast, the SNIa are not easy to assign to groups in many cases. Depending on (1) whether we fit galaxies individually or groups, (2) whether we use the ``standard candle'' model for SNIa or the ``light curve decline'' relation, and (3) whether we include or exclude NGC~5253 for which we regard our stellar population calibration as unknown, we get values for the SBF zero point as low as 13.64 and as high as 14.01 (Table 4). Averaging the two methods and again estimating uncertainties from rms divided by $\sqrt{N-1}$, we find $13.96\pm0.17$ for groups and $13.75\pm0.14$ for galaxies. Because these differ from the TF and \ifmmode D_n{-}\sigma\else$ D_n{-}\sigma$\fi\ by $2.0\;\sigma$ and $1.0\;\sigma$ respectively, the discrepancy may not be statistically significant. It is possible that there are systematic errors in the tie to the distant Hubble flow for TF and \ifmmode D_n{-}\sigma\else$ D_n{-}\sigma$\fi, whereas the SNIa appear to be wonderfully consistent with the large scale Hubble flow. On the other hand, the nearby SNIa do not agree with SBF or Cepheids as well as one might hope from the scatter against the Hubble flow, which makes one worry about the systematics with SN1a. For example, the SN1a distances predicted for the Fornax clusters are significantly larger than the very recent Cepheid measurement of the distance to the Fornax cluster (Silbermann et al. 1996b). SNII appear to agree pretty well with SBF and Cepheids, and there should eventually be enough of them to tie very well to the large scale Hubble flow. In subsequent papers we will present the direct tie between SBF and the Hubble flow, both from ground-based observations as well as HST observations beyond 5000~km/s, but at present we depend on these other estimators to tie to the Hubble flow. It is therefore with some trepidation that we offer a value for $H_0$. We have a calibration for \ifmmode\overline{M}_I\else$\overline{M}_I$\fi; we have several calibrations for \ifmmode\overline{m}_I\else$\overline{m}_I$\fi\ in terms of $5\log d\hbox{(km/s)}$; and of course $\ifmmode(m{-}M)\else$(m{-}M)$\fi = 5\log d\hbox{(km/s)} + 25 - 5\log H_0$. If we use the TF and \ifmmode D_n{-}\sigma\else$ D_n{-}\sigma$\fi\ calibration of SBF we get $H_0 = 86$~\hbox{km/s/Mpc}. Examining groups and averaging the ``standard candle'' and the ``$\Delta m_{15}$'' assumptions about SNIa gives us $H_0 = 72$~\hbox{km/s/Mpc}. If we compare galaxies directly without resorting to group membership, but leave out NGC~5253, we get an average $H_0 = 80$~\hbox{km/s/Mpc}. We suspect that there is more to the SNIa story than is currently understood, so we therefore prefer not to use it to the exclusion of all other distance estimators. The range we find for $H_0$ is $$ H_0 = 72 - 86\;\hbox{km/s/Mpc,}$$ and our best guess at this point is derived by averaging the ties to the Hubble flow from TF, \ifmmode D_n{-}\sigma\else$ D_n{-}\sigma$\fi, SNIa (both methods) in groups and SNIa (both methods) galaxy by galaxy. This weights the SNIa slightly more heavily than TF and \ifmmode D_n{-}\sigma\else$ D_n{-}\sigma$\fi\ and gives a zero point of 13.72 which translates to $$ H_0 = 81\pm6\;\hbox{km/s/Mpc.}$$ The final error term includes a contribution of 0.07 magnitude from the disagreement between the Cepheid and theory zero points (which we hope is indicative of the true accuracy of our calibrations), and an allowance of 0.13 magnitude for the uncertainty in the tie to the distant Hubble flow (judged from the scatter among the various methods). In order to facilitate comparisons with SBF distances, we offer the SBF distance to 12 nearby groups in Table 5. The relative distances are completely independent of any other distance estimator, and the zero point uses our Cepheid-based calibration. As we finish our reductions and analysis, the remainder of the group and individual galaxy distances will be published. \section{Summary and Conclusions} We have described the observational sample which comprises the SBF Survey of Galaxy Distances. The survey was conducted over numerous observing runs spanning a period of nearly seven years. The photometry of the sample has been brought into internal consistency by applying small systematic corrections to the photometric zero points of the individual runs. Based on comparisons between overlapping galaxy observations, we find that our error estimates for $(V-I)$ and $\overline m_I$ are reliable, after correction for the photometry offsets. From our measurements of \ifmmode\overline{m}_I\else$\overline{m}_I$\fi\ within galaxy groups, we conclude that \ifmmode\overline{M}_I\else$\overline{M}_I$\fi\ is well described by a linear function of \ifmmode(V{-}I)_0\else$(V{-}I)_0$\fi. Comparison of our relative distances with Cepheid distances to these groups indicates that this linear relationship is universal and yields the zero point calibration for the SBF method. This calibration is applicable to galaxies that are in the color range $1.0 < \ifmmode(V{-}I)_0\else$(V{-}I)_0$\fi < 1.3$ and which have not experienced recent bursts of star formation. Any intrinsic, or ``cosmic,'' scatter about this relation is small, of order 0.05~mag. Owing to many more data and improved photometry, this new calibration differs in its zero point by 0.35 mag from the earlier one of Tonry (1991), but is much closer to Worthey's (1993) theoretical zero point, differing by just 0.07~mag. We take this close agreement to be an independent confirmation of the Cepheid distance scale. An extensive set of comparisons between our SBF distances and those estimated using other methods provides still further evidence for the universality of the \ifmmode\overline{M}_I\else$\overline{M}_I$\fi--\ifmmode(V{-}I)_0\else$(V{-}I)_0$\fi\ relation. We find that the various methods are all generally quite reliable, apart from occasional outliers which serve to inflate the $\chi^2$ values for the comparisons. Coupled with our distance zero point, our comparisons with methods tied to the distant Hubble flow yield values of $H_0$ in the range 72--86 km/s/Mpc. The comparison with SNIa suggests values between 72 and 80, and \ifmmode D_n{-}\sigma\else$ D_n{-}\sigma$\fi\ and TF call for values around 86. Thus, the controversy over $H_0$ continues, but the famous ``factor of two'' is now a factor of 20 percent. Although the SBF Survey is still a work in progress, it is near enough to completion that the calibration presented in this paper should not change in any significant way. Future papers in this series will use the SBF survey distances to address such issues as the velocity field of the Local Supercluster and a direct determination of $H_0$, bulk flows, the Great Attractor, and the specific details of our SBF analysis method, including comprehensive listings of our \ifmmode(V{-}I)_0\else$(V{-}I)_0$\fi\ and distance measurements for individual galaxies. \acknowledgments We would like to thank many people for collecting SBF and photometry observations for us: Bob Barr, Andre Fletcher, Xiaohui Hui, Gerry Luppino, Mark Metzger, Chris Moore, and Paul Schechter. This research was supported by NSF grant AST-9401519, and AD acknowledges the support of the National Science Foundation through grant AST-9220052. \clearpage \begin{deluxetable}{llllllllllll} \small \tablecaption{Observing Runs.\label{tbl1}} \tablewidth{0pt} \tablehead{ \colhead{Run} & \colhead{Telescope} & \colhead{CCD} & \colhead{$^{\prime\prime}/p$} & \colhead{$m_{1V}$} & \colhead{$A_V$} & \colhead{$C_V$} & \colhead{$\Delta V$}& \colhead{$m_{1I}$} & \colhead{$A_I$} & \colhead{$C_I$} & \colhead{$\Delta I$} } \startdata K0389 & KPNO4m & TI-2 & 0.299 & 26.15 & 0.150 & \llap{$-$}0.070 & 0.014 & 25.42 & 0.070 & 0.000 & 0.012 \nl M1189 & MDM2.4m& ACIS & 0.465 & 23.47 & 0.179 & 0.013 & 0.000 & 22.44 & 0.065 & 0.000 & 0.000 \nl C0990 & CTIO4m & TI & 0.299 & 26.23 & 0.160 & 0.0 & \llap{$-$}0.026 & 25.29 & 0.080 & 0.0 &\llap{$-$}0.003 \nl K0990 & KPNO4m & TI-2 & 0.299 & 26.26 & 0.160 & 0.0 & 0.019 & 25.39 & 0.080 & 0.0 & 0.045 \nl H0291 & CFH3.6m& SAIC & 0.131 & 24.86 & 0.089 & 0.0 & \llap{$-$}0.016 & 24.62 & 0.033 & 0.0 & 0.029 \nl L0391 & LCO2.4m& Tek & 0.229 & 24.92 & 0.15 & 0.0 & 0.024 & 24.62 & 0.07 & 0.0 & 0.030 \nl C0491 & CTIO4m & Tek1 & 0.472 & 26.06 & 0.16 & 0.0 & 0.069 & 26.02 & 0.11 & 0.0 & 0.061 \nl K0691 & KPNO4m & TI-2 & 0.300 & 25.97 & 0.155 & 0.0 & \llap{$-$}0.002 & 25.36 & 0.06 & 0.0 & 0.019 \nl C1091 & CTIO4m & Tek2 & 0.472 & 26.21 & 0.45 & \llap{$-$}0.007 & 0.019 & 26.08 & 0.3 & 0.025 & 0.040 \nl H1091 & CFH3.6m& SAIC & 0.131 & & & & 0.000 & 24.62 & 0.07 & 0.0 & 0.000 \nl L1191 & LCO2.4m& Tek & 0.229 & 24.92 & 0.15 & 0.0 & 0.034 & 24.62 & 0.07 & 0.0 &\llap{$-$}0.014 \nl M1191 & MDM2.4m& ACIS & 0.257 & 25.11 & 0.205 & 0.0 & \llap{$-$}0.025 & 24.53 & 0.102 & 0.035 & 0.007 \nl C0492 & CTIO4m & Tek2 & 0.472 & 26.05 & 0.220 & 0.005 & 0.014 & 25.93 & 0.145 & 0.030 &\llap{$-$}0.020 \nl M0492 & MDM2.4m& Lorl & 0.343 & 24.69 & 0.33 & 0.000 & 0.010 & 24.84 & 0.20 & 0.045 & 0.007 \nl H0592 & CFH3.6m& STIS & 0.152 & 25.91 & 0.210 & 0.0 & \llap{$-$}0.010 & 25.60 & 0.110 & 0.0 &\llap{$-$}0.038 \nl M0892 & MDM2.4m& Lorl & 0.343 & 24.64 & 0.254 & 0.000 & 0.000 & 24.74 & 0.145 & 0.045 & 0.000 \nl L1092 & LCO2.4m& Tek & 0.229 & 24.92 & 0.15 & 0.0 & 0.000 & 24.62 & 0.07 & 0.0 & 0.000 \nl M1092 & MDM2.4m& Lorl & 0.343 & 24.68 & 0.32 & 0.000 & 0.010 & 24.82 & 0.22 & 0.046 & 0.029 \nl L0493 & LCO2.4m& Tek & 0.229 & 24.92 & 0.15 & 0.0 & 0.027 & 24.62 & 0.07 & 0.0 &\llap{$-$}0.070 \nl M0493 & MDM2.4m& Lorl & 0.343 & 24.67 & 0.21 & 0.022 & \llap{$-$}0.004 & 24.35 & 0.13 & 0.030 & 0.014 \nl M0493 & MDM2.4m& Tek & 0.275 & 25.32 & 0.24 & 0.005 & \llap{$-$}0.004 & 24.60 & 0.11 & 0.012 & 0.014 \nl M0593 & MDM2.4m& Lorl & 0.343 & 24.70 & 0.22 & 0.022 & \llap{$-$}0.033 & 24.35 & 0.138 & 0.030 &\llap{$-$}0.009 \nl M0593 & MDM2.4m& Tek & 0.275 & 25.32 & 0.198 & 0.025 & \llap{$-$}0.033 & 24.74 & 0.134 & 0.030 &\llap{$-$}0.009 \nl M0893 & MDM2.4m& Lorl & 0.343 & 24.82 & 0.19 & 0.012 & \llap{$-$}0.021 & 24.54 & 0.10 & 0.025 &\llap{$-$}0.006 \nl M0294 & MDM2.4m& Tek & 0.275 & 25.11 & 0.150 & \llap{$-$}0.017 & \llap{$-$}0.040 & 24.84 & 0.058 & 0.015 &\llap{$-$}0.074 \nl L0394 & LCO2.4m& Tek & 0.229 & 24.92 & 0.15 & 0.0 & 0.026 & 24.62 & 0.07 & 0.0 &\llap{$-$}0.030 \nl L0994 & LCO2.4m& Tek & 0.229 & 24.92 & 0.15 & 0.0 & 0.016 & 24.62 & 0.07 & 0.0 & 0.022 \nl G0395 & MDM1.3m& Lorl & 0.637 & 23.22 & 0.15 & 0.019 & \llap{$-$}0.029 & 22.84 & 0.064 & 0.026 &\llap{$-$}0.007 \nl G0495 & MDM1.3m& Lorl & 0.637 & 23.15 & 0.15 & 0.015 & \llap{$-$}0.041 & 22.83 & 0.064 & 0.010 &\llap{$-$}0.012 \nl G0495 & MDM1.3m& STIS & 0.445 & 23.11 & 0.15 & 0.015 & \llap{$-$}0.041 & 22.65 & 0.100 & 0.010 &\llap{$-$}0.012 \nl G0695 & MDM1.3m& Lorl & 0.637 & 23.13 & 0.15 & 0.042 & \llap{$-$}0.015 & 22.79 & 0.061 & 0.026 & 0.000 \nl G0995 & MDM1.3m& Lorl & 0.637 & 22.97 & 0.14 & 0.014 & \llap{$-$}0.009 & 22.81 & 0.05 & 0.026 &\llap{$-$}0.015 \nl G1095 & MDM1.3m& STIS & 0.445 & 22.98 & 0.15 & 0.005 & \llap{$-$}0.015 & 22.70 & 0.06 & 0.008 &\llap{$-$}0.015 \nl M0295 & MDM2.4m& Tek & 0.275 & 25.11 & 0.150 & \llap{$-$}0.014 & \llap{$-$}0.040 & 24.84 & 0.058 & 0.012 &\llap{$-$}0.074 \nl M0395 & MDM2.4m& Tek & 0.275 & 25.11 & 0.150 & \llap{$-$}0.017 & \llap{$-$}0.040 & 24.84 & 0.058 & 0.015 &\llap{$-$}0.074 \nl L0495 & LCO2.4m& Tek & 0.229 & 24.92 & 0.15 & 0.0 & 0.008 & 24.62 & 0.07 & 0.0 &\llap{$-$}0.027 \nl L1095 & LCO2.4m& Tek & 0.229 & 24.92 & 0.15 & 0.0 & 0.000 & 24.62 & 0.07 & 0.0 & 0.000 \nl M1295 & MDM2.4m& Tek & 0.275 & 25.11 & 0.150 & \llap{$-$}0.017 & \llap{$-$}0.040 & 24.84 & 0.058 & 0.015 &\llap{$-$}0.074 \nl M0196 & MDM2.4m& Tek & 0.275 & 25.11 & 0.150 & \llap{$-$}0.017 & \llap{$-$}0.040 & 24.84 & 0.058 & 0.015 &\llap{$-$}0.074 \nl M0396 & MDM2.4m& Tek & 0.275 & 25.11 & 0.150 & \llap{$-$}0.017 & \llap{$-$}0.040 & 24.84 & 0.058 & 0.015 &\llap{$-$}0.074 \nl \enddata \tablecomments{Columns: Run name, telescope, detector, plate scale (\arcsec/pixel), photometric zero point, extinction, color term, and run offset for the $V$ band then $I$ band.} \end{deluxetable} \begin{deluxetable}{lrrrrrr} \tablecaption{Landolt Fields.\label{tbl2}} \tablewidth{0pt} \tablehead{ \colhead{Field} & \colhead{RA} & \colhead{Dec} & \colhead{$V_{min}$} & \colhead{$V_{max}$} & \colhead{$\ifmmode(V{-}I)\else$(V{-}I)$\fi_{min}$} & \colhead{$\ifmmode(V{-}I)\else$(V{-}I)$\fi_{max}$} } \startdata SA92-250 & 00 54 41 & $+$00 41 11 & 14.09 & 15.35 & 0.67 & 1.34 \nl SA95-190 & 03 53 16 & $+$00 16 25 & 12.63 & 14.34 & 0.42 & 1.37 \nl SA95-275 & 03 54 40 & $+$00 27 24 & 12.17 & 14.12 & 1.40 & 2.27 \nl SA98-650 & 06 52 11 & $-$00 19 23 & 11.93 & 13.75 & 0.17 & 2.09 \nl Rubin-149 & 07 24 13 & $-$00 31 58 & 11.48 & 13.87 & $-$0.11 & 1.13 \nl PG0918+029 & 09 21 36 & $+$02 47 03 & 12.27 & 14.49 & $-$0.29 & 1.11 \nl PG1323$-$085 & 13 25 44 & $-$08 49 16 & 12.08 & 14.00 & $-$0.13 & 0.83 \nl PG1633+099 & 16 35 29 & $+$09 46 54 & 12.97 & 15.27 & $-$0.21 & 1.14 \nl SA110-232 & 18 40 50 & $+$00 01 51 & 12.52 & 14.28 & 0.89 & 2.36 \nl SA110-503 & 18 43 05 & $+$00 29 10 & 11.31 & 14.20 & 0.65 & 2.63 \nl Markarian-A & 20 43 59 & $-$10 47 42 & 13.26 & 14.82 & $-$0.24 & 1.10 \nl \enddata \tablecomments{Columns: Field name, J2000 coordinates, $V$ magnitude of the brightest and faintest star, and the \ifmmode(V{-}I)\else$(V{-}I)$\fi\ colors of the bluest and reddest star.} \end{deluxetable} \begin{deluxetable}{llrrrrrrr} \tablecaption{Nearby SBF Groups.\label{tbl3}} \tablewidth{0pt} \tablehead{ \colhead{Group} & \colhead{Example} & \colhead{RA} & \colhead{Dec} & \colhead{rad} & \colhead{$v_{ave}$} & \colhead{$v_{min}$}& \colhead{$v_{max}$} & \colhead{7S\#} } \startdata LocalGroup & N0224 & 10.0 & 41.0 & 5 & \llap{$-$}300 & \llap{$-$}500 & \llap{$-$}100 & 282 \nl Cetus & N0636 & 24.2 & \llap{$-$}7.8 & 10 & 1800 & 1500 & 2000 & 26 \nl N1023 & N1023 & 37.0 & 35.0 & 9 & 650 & 500 & 1000 & \nl N1199 & N1199 & 45.3 & \llap{$-$}15.8 & 2 & 2700 & 2500 & 3000 & 29 \nl Eridanus & N1407 & 53.0 & \llap{$-$}21.0 & 6 & 1700 & 500 & 2300 & 32 \nl Fornax & N1399 & 54.1 & \llap{$-$}35.6 & 6 & 1400 & 500 & 2100 & 31 \nl Dorado & N1549 & 63.7 & \llap{$-$}55.7 & 5 & 1300 & 700 & 1700 & 211 \nl N1700 & N1700 & 72.2 & \llap{$-$}3.5 & 3 & 4230 & 3600 & 4500 & 100 \nl N2768 & N2768 & 136.9 & 60.2 & 4 & 1360 & 1100 & 1700 & 215 \nl M81 & N3031 & 147.9 & 69.3 & 8 & \llap{$-$}40 & \llap{$-$}200 & 400 & \nl N3115 & N3115 & 150.7 & \llap{$-$}7.5 & 8 & 700 & 100 & 900 & \nl LeoIII & N3193 & 153.9 & 22.1 & 3 & 1400 & 1000 & 1700 & 45 \nl LeoI & N3379 & 161.3 & 12.8 & 2 & 900 & 500 & 1200 & 57 \nl LeoII & N3607 & 168.6 & 18.3 & 3 & 950 & 650 & 1500 & 48 \nl N3640 & N3640 & 169.6 & 3.5 & 2 & 1300 & 1200 & 1800 & 50 \nl UMa & N3928 & 180.0 & 47.0 & 8 & 900 & 700 & 1100 & \nl N4125 & N4125 & 181.4 & 65.5 & 3 & 1300 & 1000 & 1700 & 54 \nl VirgoW & N4261 & 184.2 & 6.1 & 2 & 2200 & 2000 & 2800 & \nl ComaI & N4278 & 184.4 & 29.6 & 3 & 1000 & 200 & 1400 & 55 \nl CVn & N4258 & 185.0 & 44.0 & 7 & 500 & 400 & 600 & \nl N4386 & N4386 & 185.6 & 75.8 & 5 & 1650 & 1500 & 2100 & 98 \nl N4373 & N4373 & 185.7 & \llap{$-$}39.5 & 2 & 3400 & 2500 & 3800 & 35 \nl Virgo & N4486 & 187.1 & 12.7 & 10 & 1150 & \llap{$-$}300 & 2000 & 56 \nl ComaII & N4494 & 187.2 & 26.1 & 5 & 1350 & 1200 & 1400 & 235 \nl N4594 & N4594 & 189.4 & \llap{$-$}11.4 & 5 & 1100 & 900 & 1200 & \nl M51 & N5194 & 200.0 & 45.0 & 4 & 480 & 380 & 580 & \nl Centaurus & N4696 & 191.5 & \llap{$-$}41.0 & 3 & 3000 & 2000 & 5000 & 58 \nl CenA & N5128 & 200.0 & \llap{$-$}39.0 & 15 & 550 & 200 & 600 & 226 \nl N5322 & N5322 & 212.5 & 57.0 & 6 & 2000 & 1600 & 2400 & 245 \nl N5638 & N5638 & 216.0 & 3.5 & 3 & 1650 & 1400 & 1900 & 68 \nl N5846 & N5846 & 226.0 & 1.8 & 2 & 1700 & 1200 & 2200 & 70 \nl N5898 & N5898 & 228.8 & \llap{$-$}23.9 & 2 & 2100 & 2000 & 2700 & 71 \nl N6684 & N6684 & 281.0 & \llap{$-$}65.2 & 10 & 850 & 500 & 1200 & 78 \nl N7144 & N7144 & 327.4 & \llap{$-$}48.5 & 6 & 1900 & 1500 & 2000 & 84 \nl N7180 & N7180 & 329.9 & \llap{$-$}20.8 & 10 & 1500 & 1300 & 1900 & 265 \nl N7331 & N7457 & 338.7 & 34.2 & 9 & 800 & 800 & 1100 & \nl Grus & I1459 & 343.6 & \llap{$-$}36.7 & 5 & 1600 & 1400 & 2300 & 231 \nl \enddata \tablecomments{Columns: Group name, sample member, RA and Dec (B1950), group radius (deg), mean heliocentric velocity, minimum and maximum velocities for inclusion in the group, and group number from Faber et al. (1989)} \end{deluxetable} \begin{deluxetable}{lccrrrrrl} \tablecaption{Distance Comparisons.\label{tbl4}} \tablewidth{0pt} \tablehead{ \colhead{Estimator} & \colhead{Grp/gxy} & \colhead{Distance} & \colhead{N} & \colhead{$\langle\overline{m}_I^0\rangle-d$} & \colhead{$\pm$} & \colhead{rms} & \colhead{$\chi^2/N$} & \colhead{Comments} } \startdata Cepheid & Grp & (m-M) & 7 & \llap{$-$}1.74 & 0.05 & 0.16 & 0.6 & \nl Cepheid & gxy & (m-M) & 5 & \llap{$-$}1.75 & 0.06 & 0.33 & 3.4 & \nl Cepheid & gxy & (m-M) & 4 & \llap{$-$}1.82 & 0.07 & 0.12 & 0.3 & less N5253 \nl PNLF & Grp & (m-M) & 12 & \llap{$-$}1.63 & 0.02 & 0.33 & 7.5 & \nl PNLF & Grp & (m-M) & 10 & \llap{$-$}1.69 & 0.03 & 0.20 & 2.2 & less ComaI/II \nl SNII & Grp & (m-M) & 5 & \llap{$-$}1.80 & 0.12 & 0.36 & 1.4 & \nl SNII & Grp & (m-M) & 4 & \llap{$-$}1.76 & 0.12 & 0.22 & 1.1 & less N7331 \nl TF & Grp & (m-M) & 26 & \llap{$-$}1.69 & 0.03 & 0.41 & 2.1 & \nl TF (MkII) & Grp & 5logd & 29 & 13.55 & 0.08 & 0.59 & 2.1 & \nl Dn-sigma & Grp & 5logd & 28 & 13.64 & 0.05 & 0.44 & 1.9 & \nl SNIa ($M_{max}$)& Grp & 5logd & 6 & 13.92 & 0.08 & 0.38 & 3.6 & \nl SNIa ($\Delta m_{15}$)& Grp & 5logd & 6 & 14.01 & 0.08 & 0.40 & 3.6 & \nl SNIa ($M_{max}$)& gxy & 5logd & 5 & 13.86 & 0.12 & 0.54 & 4.9 & \nl SNIa ($\Delta m_{15}$)& gxy & 5logd & 5 & 14.01 & 0.12 & 0.43 & 3.2 & \nl SNIa ($M_{max}$)& gxy & 5logd & 4 & 13.64 & 0.13 & 0.22 & 1.0 & less N5253 \nl SNIa ($\Delta m_{15}$)& gxy & 5logd & 4 & 13.87 & 0.13 & 0.30 & 1.8 & less N5253 \nl Theory & & & & \llap{$-$}1.81 & & 0.11 & & \nl \enddata \tablecomments{Columns: Name of the estimator, comparison by group or by galaxy, estimator's zero point based on Mpc \ifmmode(m{-}M)\else$(m{-}M)$\fi\ or Hubble flow (5logd~km/s), number of comparison points, mean difference between SBF and the estimator, expected error in this mean based on error estimates, rms scatter in the comparison, $\chi^2/N$, and comments.} \end{deluxetable} \begin{deluxetable}{llrrrrrrrl} \tablecaption{SBF Distances to Groups.\label{tbl5}} \tablewidth{0pt} \tablehead{ \colhead{Group} & \colhead{Example} & \colhead{RA} & \colhead{Dec} & \colhead{$v_{ave}$} & \colhead{$N$} & \colhead{\ifmmode(m{-}M)\else$(m{-}M)$\fi} & \colhead{$\pm$} & \colhead{$d$} & $\pm$ } \startdata LocalGrp & N0224 & 10.0 & 41.0 & \llap{$-$}300 & 2 & 24.43 & 0.08 & \phn0.77 & 0.03 \nl M81 & N3031 & 147.9 & 69.3 & \llap{$-$}40 & 2 & 27.78 & 0.08 & \phn3.6 & 0.2 \nl CenA & N5128 & 200.0 & \llap{$-$}39.0 & 550 & 3 & 28.03 & 0.10 & \phn4.0 & 0.2 \nl N1023 & N1023 & 37.0 & 35.0 & 650 & 4 & 29.91 & 0.09 & \phn9.6 & 0.4 \nl LeoI & N3379 & 161.3 & 12.8 & 900 & 5 & 30.14 & 0.06 & 10.7 & 0.3 \nl N7331 & N7331 & 338.7 & 34.2 & 800 & 2 & 30.39 & 0.10 & 12.0 & 0.6 \nl UMa & N3928 & 180.0 & 47.0 & 900 & 5 & 30.76 & 0.09 & 14.2 & 0.6 \nl ComaI & N4278 & 184.4 & 29.6 & 1000 & 3 & 30.95 & 0.08 & 15.5 & 0.6 \nl ComaII & N4494 & 187.2 & 26.1 & 1350 & 3 & 31.01 & 0.08 & 15.9 & 0.6 \nl Virgo & N4486 & 187.1 & 12.7 & 1150 & 27 & 31.03 & 0.05 & 16.1 & 0.4 \nl Dorado & N1549 & 63.7 & \llap{$-$}55.7 & 1300 & 6 & 31.04 & 0.06 & 16.1 & 0.5 \nl Fornax & N1399 & 54.1 & \llap{$-$}35.6 & 1400 & 26 & 31.23 & 0.06 & 17.6 & 0.5 \nl \enddata \tablecomments{Columns: Group name, sample member, RA and Dec (B1950), mean heliocentric velocity, number of SBF distances, SBF distance modulus and error, and SBF distance (Mpc) and error.} \end{deluxetable} \clearpage
proofpile-arXiv_065-420
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{INTRODUCTION} This talk reviews the status of two distinct though related subjects: our understanding of chiral extrapolations, and results for weak matrix elements involving light (i.e. $u$, $d$ and $s$) quarks. The major connection between these subjects is that understanding chiral extrapolations allows us to reduce, or at least estimate, the errors in matrix elements. Indeed, in a number of matrix elements, the dominant errors are those due to chiral extrapolation and to the use of the quenched approximation. I will argue that an understanding of chiral extrapolations gives us a handle on both of these errors. While understanding errors in detail is a sign of a maturing field, we are ultimately interested in the results for the matrix elements themselves. The phenomenological implications of these results were emphasized in previous reviews \cite{martinelli94,soni95}. Here I only note that it is important to calculate matrix elements for which we know the answer, e.g. $f_\pi/M_\rho$ and $f_K/f_\pi$, in order to convince ourselves, and others, that our predictions for unknown matrix elements are reliable. The reliability of a result for $f_D$, for example, will be gauged in part by how well we can calculate $f_\pi$. And it would be a real coup if we were able to show in detail that QCD indeed explains the $\Delta I=1/2$ rule in $K\to\pi\pi$ decays. But to me the most interesting part of the enterprise is the possibility of using the lattice results to calculate quantities which allow us to test the Standard Model. In this category, the light-quark matrix element with which we have had the most success is $B_K$. The lattice result is already used by phenomenological analyses which attempt to determine the CP violation in the CKM matrix from the experimental number for $\epsilon$. I describe below the latest twists in the saga of the lattice result for $B_K$. What we would like to do is extend this success to the raft of $B$-parameters which are needed to predict $\epsilon'/\epsilon$. There has been some progress this year on the contributions from electromagnetic penguins, but we have made no headway towards calculating strong penguins. I note parenthetically that another input into the prediction of $\epsilon'/\epsilon$ is the strange quark mass. The recent work of Gupta and Bhattacharya \cite{rajanmq96} and the Fermilab group \cite{mackenzie96} suggest a value of $m_s$ considerably smaller than the accepted phenomenological estimates, which will substantially increase the prediction for $\epsilon'/\epsilon$. Much of the preceding could have been written in 1989, when I gave the review talk on weak matrix elements \cite{sharpe89}. How has the field progressed since then? I see considerable advances in two areas. First, the entire field of weak matrix elements involving heavy-light hadrons has blossomed. 1989 was early days in our calculation of the simplest such matrix elements, $f_D$ and $f_B$. In 1996 a plethora of quantities are being calculated, and the subject deserves its own plenary talk \cite{flynn96}. Second, while the subject of my talks then and now is similar, there has been enormous progress in understanding and reducing systematic errors. Thus, whereas in 1989 I noted the possibility of using chiral loops to estimate quenching errors, we now have a technology (quenched chiral perturbation theory---QChPT) which allows us to make these estimates. We have learned that the quenched approximation (QQCD) is most probably singular in the chiral limit, and there is a growing body of numerical evidence showing this, although the case is not closed. The reduction in statistical errors has allowed us to go beyond simple linear extrapolations in light quark masses, and thus begin to test the predictions of QChPT. The increase in computer power has allowed us to study systematically the dependence of matrix elements on the lattice spacing, $a$. We have learned how to get more reliable estimates using lattice perturbation theory \cite{lepagemackenzie}. And, finally, we have begun to use non-perturbative matching of lattice and continuum operators, as discussed here by Rossi \cite{rossi96}. The body of this talk is divided into two parts. In the first, Secs. \ref{sec:whychi}-\ref{sec:querr}, I focus on chiral extrapolations: why we need them, how we calculate their expected form, the evidence for chiral loops, and how we can use them to estimate quenching errors. In the second part, Secs. \ref{sec:decayc}-\ref{sec:otherme}, I give an update on results for weak matrix elements. I discuss $f_\pi/M_\rho$, $f_K/f_\pi$, $B_K$ and a few related $B$-parameters. Results for structure functions have been reviewed here by G\"ockeler \cite{gockeler96}. There has been little progress on semi-leptonic form factors, nor on flavor singlet matrix elements and scattering lengths, since last years talks by Simone \cite{simone95} and Okawa \cite{okawa95}. \section{WHY DO WE NEED QChPT?} \label{sec:whychi} Until recently linear chiral extrapolations (i.e. $\alpha + \beta m_q$) have sufficed for most quantities. This is no longer true. This change has come about for two reasons. First, smaller statistical errors (and to some extent the use of a larger range of quark masses) have exposed the inadequacies of linear fits, as already stressed here by Gottlieb \cite{gottlieb96}. An example, taken from Ref. \cite{ourspect96} is shown in Fig. \ref{fig:NDelta}. The range of quark masses is $m_s/3- 2 m_s$, typical of that in present calculations. While $M_\Delta$ is adequately fit by a straight line, there is definite, though small, curvature in $M_N$. The curves are the result of a fit using QChPT \cite{sharpebary}, to which I return below. \begin{figure}[tb] \vspace{-0.1truein} \centerline{\psfig{file=fig1.ps,height=3.0truein}} \vspace{-0.6truein} \caption{$M_N$ and $M_\Delta$ versus $M_\pi^2$, with quenched Wilson fermions, at $\beta=6$, on $32^3\times64$ lattices. The vertical band is the range of estimates for $m_s$.} \vspace{-0.2truein} \label{fig:NDelta} \end{figure} The second demonstration of the failure of linear extrapolations has come from studying the octet baryon mass splittings \cite{ourspect96}. The new feature here is the consideration of baryons composed of non-degenerate quarks. I show the results for $(M_\Sigma-M_\Lambda)/(m_s-m_u)$ (with $m_u=m_d$) in Fig. \ref{fig:sigdel}. If baryons masses were linear functions of $m_s$ and $m_u$ then the data would lie on a horizontal line. Instead the results vary by a factor of four. This is a glaring deviation from linear behavior, in contrast to the subtle effect shown in Fig. \ref{fig:NDelta}. \begin{figure}[tb] \vspace{-0.1truein} \centerline{\psfig{file=fig2.ps,height=3.0truein}} \vspace{-0.6truein} \caption{Results for $(M_\Sigma-M_\Lambda)/(m_s-m_u)$ (diamonds). The ``burst'' is the physical result using the linear fit shown. Crosses are from a global chiral fit.} \vspace{-0.2truein} \label{fig:sigdel} \end{figure} Once there is evidence of non-linearity, we need to know the appropriate functional form with which to extrapolate to the chiral limit. This is where (Q)ChPT comes in. In the second example, the prediction is \cite{LS,sharpebary} \begin{eqnarray} \lefteqn{{M_\Sigma - M_\Lambda \over m_s-m_u} \approx} \nonumber\\ && - {8 D\over3} + d_1 {M_K^3-M_\pi^3 \over m_s-m_u} + d'_1 {M_{ss}^3-M_\pi^3 \over m_s-m_u} \nonumber \\ && + e_1 (m_u + m_d + m_s) + e'_1 (m_u + m_s) \,, \label{eq:chptsiglam} \end{eqnarray} where $M_{ss}$ is the mass of the quenched $\bar ss$ meson.\footnote{% I have omitted terms which, though more singular in the chiral limit, are expected to be numerically small for the range of $m_q$ under study.} QChPT also implies that $|d'_1| < |d_1|$, which is why in Fig. \ref{fig:sigdel} I have plotted the data versus $(M_K^3-M_\pi^3)/(m_s-m_u)$. The good news is that the points do lie approximately on a single curve---which would not be true for a poor choice of y-axis. The bad news is that the fit to the the $d_1$ term and a constant is not that good. This example illustrates the benefits which can accrue if one knows the chiral expansion of the quantity under study. First, the data collapses onto a single curve, allowing an extrapolation to the physical quark masses. And, second, the theoretical input reduces the length of the extrapolation. In Fig. \ref{fig:sigdel}, for example, to reach the physical point requires an extrapolation by less than a factor of 2. This is much smaller than the ratio of the lightest quark mass to the physical value of $(m_u+m_d)/2$---a factor of roughly 8. In fact, in the present example, the data are not yet accurate enough to distinguish between the $d$ and $e$ terms in Eq. \ref{eq:chptsiglam}. A global fit of the QChPT prediction to this and other mass differences (the results of which are shown in the Figure) implies that both types of term are present \cite{sharpebary}. \section CHIRAL PERTURBATION THEORY} \label{sec:qchpt} This brings me to a summary of QChPT. In QCD, chiral perturbation theory predicts the form of the chiral expansion for quantities involving one or more light quarks. The expansion involves terms analytic in $m_q$ and the external momenta, and non-analytic terms due to pion loops.\footnote{``Pion'' here refers to any of the pseudo-Goldstone bosons.} The analytic terms are restricted in form, though not in magnitude, by chiral symmetry, while the non-analytic terms are completely predicted given the analytic terms. The same type of expansions can be developed in quenched QCD using QChPT. The method was worked out in Refs. \cite{morel,sharpestcoup,BGI,sharpechbk}, with the theoretically best motivated formulation being that of Bernard and Golterman \cite{BGI}. Their method gives a precise meaning to the quark-flow diagrams which I use below. They have extended it also to partially quenched theories (those having both valence and sea quarks but with different masses) \cite{BGPQ}. Results are available for pion properties and the condensate \cite{BGI}, $B_K$ and related matrix elements \cite{sharpechbk}, $f_B$, $B_B$ and the Isgur-Wise function \cite{booth,zhang}, baryon masses \cite{sharpebary,LS}, and scattering lengths \cite{BGscat}. I will not describe technical details, but rather focus on the aims and major conclusions of the approach. For a more technical review see Ref. \cite{maartench}. As I see it, the major aims of QChPT are these. \\ $\bullet\ $ To predict the form of the chiral expansion, which can then be used to fit and extrapolate the data. This was the approach taken above for the baryon masses. \\ $\bullet\ $ To estimate the size of the quenching error by comparing the contribution of the pion loops in QCD and QQCD, using the coefficients of the chiral fits in QCD (from phenomenology) and in QQCD (from a fit to the lattice data). I return to such estimates in Sec. \ref{sec:querr}. I begin by describing the form of the chiral expansions, and in particular how they are affected by quenching. The largest changes are to the non-analytic terms, i.e. those due to pion loops. There are two distinct effects.\\ (1) Quenching removes loops which, at the underlying quark level, require an internal loop. This is illustrated for baryon masses in Fig. \ref{fig:quarkloops}. Diagrams of types (a) and (b) contribute in QCD, but only type (b) occurs in QQCD. These loops give rise to $M_\pi^3$ terms in the chiral expansions. Thus for baryon masses, quenching only changes the coefficient of these terms. In other quantities, e.g. $f_\pi$, they are removed entirely. \\ (2) Quenching introduces artifacts due to $\eta'$ loops---as in Fig. \ref{fig:quarkloops}(c). These are chiral loops because the $\eta'$ remains light in QQCD. Their strength is determined by the size of the ``hairpin'' vertex, and is parameterized by $\delta$ and $\alpha_\Phi$ (defined in Sec. \ref{sec:chevidence} below). \begin{figure}[tb] \centerline{\psfig{file=fig3.ps,height=2.4truein}} \vspace{-0.3truein} \caption{Quark flow diagrams for $M_{\rm bary}$.} \vspace{-0.2truein} \label{fig:quarkloops} \end{figure} The first effect of quenching is of greater practical importance, and dominates most estimates of quenching errors. It does not change the form of the chiral expansion, only the size of the terms. The second effect leads to new terms in the chiral expansion, some of which are singular in the chiral limit. If these terms are large, then one can be sure that quenching errors are large. One wants to work at large enough quark mass so that these new terms are numerically small. In practice this means keeping $m_q$ above about $m_s/4-m_s/3$. I illustrate these general comments using the results of Labrenz and I for baryon masses \cite{LS}. The form of the chiral expansion in QCD is\footnote{% There are also $M_\pi^4 \log M_\pi$ terms in both QCD and QQCD, the coefficients of which have not yet been calculated in QQCD. For the limited range of quark masses used in simulations, I expect that these terms can be adequately represented by the $M_\pi^4$ terms, whose coefficients are unknown parameters.} \begin{equation} M_{\rm bary} = M_0 + c_2 M_\pi^2 + c_3 M_\pi^3 + c_4 M_\pi^4 + \dots \end{equation} with $c_3$ predicted in terms of $g_{\pi NN}$ and $f_\pi$. In QQCD \begin{eqnarray} \lefteqn{ M_{\rm bary}^Q = M_0^Q + c_2^Q M_\pi^2 + c_3^Q M_\pi^3 + c_4^Q M_\pi^4 + } \nonumber \\ && \delta \left( c_1^Q M_\pi + c_2^Q M_\pi^2 \log M_\pi\right) + \alpha_\Phi \tilde c_3^Q M_\pi^3 + \dots \label{eq:baryqchpt} \end{eqnarray} The first line has the same form as in QCD, although the constants multiplying the analytic terms in the two theories are unrelated. $c_3^Q$ is predicted to be non-vanishing, though different from $c_3$. The second line is the contribution of $\eta'$ loops and is a quenched artifact. Note that it is the dominant correction in the chiral limit. In order to test QChPT in more detail, I have attempted to fit the expressions outlined above to the octet and decuplet baryon masses from Ref. \cite{ourspect96}. There are 48 masses, to be fit in terms of 19 underlying parameters: the octet and decuplet masses in the chiral limit, 3 constants of the form $c_2^Q$, 6 of the form $c_4^Q$, 6 pion-nucleon couplings, $\delta$ and $\alpha_\Phi$. I have found a reasonable description of the data with $\delta\approx0.1$, but the errors are too large to pin down the values of all the constants \cite{sharpebary}. Examples of the fit are shown in Figs. \ref{fig:NDelta} and \ref{fig:sigdel}. \begin{figure}[tb] \vspace{-0.1truein} \centerline{\psfig{file=fig4.ps,height=3.0truein}} \vspace{-0.6truein} \caption{Contributions to $M_N - M_0^Q$ in the global chiral fit. All quantities in lattice units. Vertical lines indicate range of estimates of $m_s$.} \vspace{-0.2truein} \label{fig:mnchfit} \end{figure} Although this is a first, rather crude, attempt at such a fit, several important lessons emerge.\\ (1) For present quark masses, one needs several terms in the chiral expansion to fit the data. This in turn requires that one have high statistics results for a number of light quark masses. \\ (2) One must check that the ``fit'' is not resulting from large cancellations between different orders. The situation for my fit is illustrated by Fig. \ref{fig:mnchfit}, where I show the different contributions to $M_N$. Note that the relative size of the terms is not determined by $M_N$, but rather by the fit as a whole. The most important point is that the $M_\pi^4$ terms are considerably smaller than those of $O(M_\pi^2)$ up to at least $m_q= m_s$. The $M_\pi^3$ terms are part of a different series and need not be smaller than those of $O(M_\pi^2)$. Similarly the $M_\pi$ terms are the leading quenched artifact, and should not be compared to the other terms. Thus the convergence is acceptable for $m_q < m_s$, though it is dubious for the highest mass point. \\ (3) The artifacts (in particular the $\delta M_\pi$ terms) can lead to unusual behavior at small $M_\pi$, as illustrated in the fit to $M_\Delta$ (Fig. \ref{fig:NDelta}). \\ (4) Since the ``$\delta$-terms'' are artifacts of quenching, and their relative contribution increases as $M_\pi\to0$, it makes more sense phenomenologically {\em to extrapolate without including them}. In other words, a better estimate of the unquenched value for $M_\Delta$ in the chiral limit can probably be obtained simply using a linear extrapolation in $M_\pi^2$. This is, however, a complicated issue which needs more thought.\\ (5) The output of the fit includes pion-nucleon couplings whose values should be compared to more direct determinations. \\ (6) Finally, the fact that a fit can be found at all gives me confidence to stick my neck out and proceed with the estimates of quenching errors in baryon masses. It should be noted, however, that a fit involving only analytic terms, including up to $M_\pi^6$, can probably not be ruled out. What of quantities other than baryon masses? In Sec. \ref{subsec:bkchfit} I discuss fits to $B_K$, another quantity in which chiral loops survive quenching. The data is consistent with the non-analytic term predicted by QChPT. Good data also exists for $M_\rho$. It shows curvature, but is consistent with either cubic or quartic terms\cite{sloan96}. What do we expect from QChPT? In QCD the chiral expansion for $M_\rho$ has the same form as for baryon masses \cite{rhochpt}. The QChPT theory calculation has not been done, but it is simple to see that form will be as for baryons, Eq.~\ref{eq:baryqchpt}, {\em except that $c_3^Q=0$}. Thus an $M_\pi^3$ term is entirely a quenched artifact---and a potential window on $\alpha_\Phi$. What of quantities involving pions, for which there is very good data? For the most part, quenching simply removes the non-analytic terms of QCD and replaces them with artifacts proportional to $\delta$. The search for these is the subject of the next section. \section{EVIDENCE FOR $\eta'$ LOOPS} \label{sec:chevidence} The credibility of QChPT rests in part on the observation of the singularities predicted in the chiral limit. If such quenched artifacts are present, then we need to study them if only to know at what quark masses to work in order to avoid them! What follows in this section is an update of the 1994 review of Gupta \cite{gupta94}. The most direct way of measuring $\delta$ is from the $\eta'$ correlator. If the quarks are degenerate, then the part of the quenched chiral Lagrangian bilinear in the $\eta'$ is \cite{BGI,sharpechbk} \begin{eqnarray} 2 {\cal L}_{\eta'} &=& \partial_\mu \eta' \partial^\mu \eta' - M_\pi^2 \eta'^2 \\ &&+ {(N_f/3)} \left( \alpha_\Phi \partial_\mu \eta' \partial^\mu \eta' - m_0^2 \eta'^2 \right) \,. \end{eqnarray} $\delta$ is defined by $\delta = m_0^2/(48 \pi^2 f_\pi^2)$. In QQCD the terms in the second line must be treated as vertices, and cannot be iterated. They contribute to the disconnected part of the $\eta'$ correlator, whereas the first line determines the connected part. Thus the $\eta'$ is degenerate with the pion, but has additional vertices. To study this, various groups have looked at the ratio of the disconnected to connected parts, at $\vec p=0$, whose predicted form at long times is \begin{equation} R(t) = t (N_f/3) (m_0^2 - \alpha_\Phi M_\pi^2)/(2 M_\pi) \,. \end{equation} I collect the results in Table \ref{tab:etapres}, converting $m_0^2$ into $\delta$ using $a$ determined from $m_\rho$. \begin{table}[tb] \caption{Results from the quenched $\eta'$ two point function. $W$ and $S$ denote Wilson and staggered fermions.} \label{tab:etapres} \begin{tabular}{ccccc} \hline Ref. & Yr. &$\delta$&$\alpha_\Phi$ &$\beta$ W/S \\ \hline JLQCD\cite{kuramashi94} &94&$0.14 (01)$ & $0.7$ & $5.7$ W \\ OSU\cite{kilcup95}&95& $0.27 (10)$ & $0.6$ & $6.0$ S \\ Rome\cite{masetti96}&96& $\approx 0.15$ & & $5.7$ W \\ OSU\cite{venkat96}& 96&$0.19(05)$ & $0.6$ & $6.0$ S \\ FNAL\cite{thacker96} &96& $< 0.02$ & $>0$ & $5.7$ W \\ \hline \end{tabular} \vspace{-0.2truein} \end{table} All groups except Ref. \cite{thacker96} report a non-zero value for $\delta$ in the range $0.1-0.3$. What they actually measure, as illustrated in Fig. \ref{fig:osuR}, is the combination $m_0^2-\alpha_\Phi M_\pi^2$, which they then extrapolate to $M_\pi=0$. I have extracted the results for $\alpha_\Phi$ from such plots. As the figure shows, there is a considerable cancellation between the $m_0$ and $\alpha_\Phi$ terms at the largest quark masses, which correspond to $M_\pi \approx 0.8\,$GeV. This may explain why Ref. \cite{thacker96} does not see a signal for $\delta$. \begin{figure}[tb] \vspace{-0.3truein} \centerline{\psfig{file=fig5.ps,height=2.4truein}} \vspace{-0.5truein} \caption{$a^2(m_0^2-\alpha_\Phi M_\pi^2)$ from the OSU group.} \vspace{-0.2truein} \label{fig:osuR} \end{figure} Clearly, further work is needed to sort out the differences between the various groups. As emphasized by Thacker \cite{thacker96}, this is mainly an issue of understanding systematic errors. In particular, contamination from excited states leads to an apparent linear rise of $R(t)$, and thus to an overestimate of $m_0^2$. Indeed, the difference between the OSU results last year and this is the use of smeared sources to reduce such contamination. This leads to a smaller $\delta$, as shown in Fig. \ref{fig:osuR}. Ref. \cite{thacker96} also find that $\delta$ decreases as the volume is increased. I want to mention also that the $\eta'$ correlator has also been studied in partially quenched theories, with $N_f=-6, -4, -2$, \cite{masetti96} and $N_f=2, 4$ \cite{venkat96}. The former work is part of the ``bermion'' program which aims to extrapolate from negative to positive $N_f$. For any non-zero $N_f$ the analysis is different than in the quenched theory, because the hairpin vertices do iterate, and lead to a shift in the $\eta'$ mass. Indeed $m_{\eta'}$ is reduced (increased) for $N_f<0$ ($>0$), and both changes are observed! This gives me more confidence in the results of these groups at $N_f=0$. The bottom line appears to be that there is relatively little dependence of $m_0^2$ on $N_f$. Other ways of obtaining $\delta$ rely on loop effects, such as that in Fig. \ref{fig:quarkloops}(c). For quenched pion masses $\eta'$ loops lead to terms which are singular in the chiral limit \cite{BGI}\footnote{% The $\alpha_\Phi$ vertex leads to terms proportional to $M_\pi^2 \log M_\pi$ which are not singular in the chiral limit, and can be represented approximately by analytic terms.} \begin{eqnarray} \lefteqn{{M_{12}^2 \over m_1 + m_2} = \mu^Q \left[ 1 - \delta \left\{\log{\widetilde M_{11}^2\over \Lambda^2} \right. \right.} \nonumber \\ && \left.\left. + {\widetilde M_{22}^2 \over \widetilde M_{22}^2 -\widetilde M_{11}^2 } \log{\widetilde M_{22}^2 \over \widetilde M_{11}^2} \right\} + c_2 (m_1 + m_2) \right] \label{eq:mpichpt} \end{eqnarray} Here $M_{ij}$ is the mass of pion composed of a quark of mass $m_i$ and antiquark of mass $m_j$, $\Lambda$ is an unknown scale, and $c_2$ an unknown constant. The tilde is relevant only to staggered fermions, and indicates that it is the mass of the flavor singlet pion, and not of the lattice pseudo-Goldstone pion, which appears. This is important because, at finite lattice spacing, $\tilde M_{ii}$ does not vanish in the chiral limit, so there is no true singularity. In his 1994 review, Gupta fit the world's data for staggered fermions at $\beta=6$ having $m_1=m_2$. I have updated his plot, including new JLQCD data \cite{yoshie96}, in Fig. \ref{fig:mpibymq}. To set the scale, note that $m_s a \approx 0.024$. The dashed line is Gupta's fit to Eq. \ref{eq:mpichpt}, giving $\delta=0.085$, while the solid line includes also an $m_q^2$ term, and gives $\delta=0.13$. These non-zero values were driven by the results from Kim and Sinclair (KS), who use quark masses as low as $0.1 m_s$ \cite{kimsinclair}, but they are now supported by the JLQCD results. \begin{figure}[tb] \vspace{-0.1truein} \centerline{\psfig{file=fig6.ps,height=3.0truein}} \vspace{-0.6truein} \caption{Chiral fit to $\log(M_\pi^2/m_q)$ at $\beta=6$. Some points have been offset for clarity.} \vspace{-0.2truein} \label{fig:mpibymq} \end{figure} Last year, Mawhinney proposed an alternative explanation for the increase visible at small $m_q$ \cite{mawhinney95}, namely an offset in the intercept of $M_\pi^2$ \begin{equation} M_{12}^2 = c_0 + \mu^Q (m_1 + m_2) + \dots \label{eq:mpimawh} \end{equation} In his model, $c_0$ is proportional to the minimum eigenvalue of the Dirac operator, and thus falls as $1/V$. This model explains the detailed structure of his results for $M_\pi^2$ and $\langle \bar\psi\psi \rangle$ at $\beta=5.7$. It also describes the data of KS, {\em except for volume dependence of $c_0$}. As the Fig. \ref{fig:mpibymq} shows, the results of KS from $24^3$ and $32^3$ lattices are consistent, whereas in Mawhinney's model the rise at small $m_q$ should be reduced in amplitude by $0.4$ on the larger lattice. The fit in Fig. \ref{fig:mpibymq} is for pions composed of degenerate quarks. One can further test QChPT by noting that Eq. \ref{eq:mpichpt} is not simply a function of the average quark mass---there is a predicted dependence on $m_1-m_2$. In Mawhinney's model, this dependence would presumably enter only through a $(m_1-m_2)^2$ term, and thus would be a weaker effect. JLQCD have extensive data from the range $\beta=5.7-6.4$, with both $m_1=m_2$ and $m_1\ne m_2$. They have fit to Eq. \ref{eq:mpichpt}, and thus obtained $\delta$ as a function of $\beta$. They find reasonable fits, with $\delta\approx 0.06$ for most $\beta$. I have several comments on their fits. First, they have used $M_{ii}$ rather than $\widetilde M_{ii}$, which leads to an underestimate of $\delta$, particularly at the smaller values of $\beta$. Second, their results for the constants, particularly $c_2$, vary rapidly with $\beta$. One would expect that all dimensionless parameters in the fit (which are no less physical than, say, $f_\pi/M_\rho$) should vary smoothly and slowly with $\beta$. This suggests to me that terms of $O(m_q^2)$ may be needed. Finally, it would be interesting to attempt a fit to the JLQCD data along the lines suggested by Mawhinney, but including a $(m_1-m_2)^2$ term. Clearly more work is needed to establish convincingly that there are chiral singularities in $M_\pi$. One should keep in mind that the effects are small, $\sim 5\%$ at the lightest $m_q$, so it is impressive that we can study them at all. Let me mention also some other complications.\\ (1) It will be hard to see the ``singularities'' with staggered fermions for $\beta<6$. This is because $\widetilde M_{ii} - M_{ii}$ grows like $a^2$ (at fixed physical quark mass). Indeed, by $\beta=5.7$ the flavor singlet pion has a mass comparable to $M_\rho$! Thus the $\eta'$ is no longer light, and its loop effects will be suppressed. In fact, the rise in $M_\pi^2/m_q$ as $m_q\to0$ for $\beta=5.7$ is very gradual\cite{gottlieb96}, and could be due to the $c_2$ term.\\ (2) It will be hard to see the singularities using Wilson fermions. This is because we do not know, {\it a priori}, where $m_q$ vanishes, and, as shown by Mawhinney, it is hard to distinguish the log divergence of QChPT from an offset in $m_q$.\\ (3) A related log divergence is predicted for $\langle\bar\psi\psi\rangle$, which has not been seen so far \cite{kimsinclair,mawhinney95}. It is not clear to me that this is a problem for QChPT, however, because it is difficult to extract the non-perturbative part of $\langle\bar\psi\psi\rangle$ from the quadratically divergent perturbative background. Two other quantities give evidence concerning $\delta$. The first uses the ratio of decay constants \begin{equation} R_{BG} = f_{12}^2/(f_{11} f_{22}) \,. \end{equation} This is designed to cancel the analytic terms proportional to $m_q$ \cite{BGI}, leaving a non-analytic term proportional to $\delta$. The latest analysis finds $\delta\approx 0.14$ \cite{guptafpi}. It is noteworthy that a good fit once again requires the inclusion of $O(m_q^2)$ terms. The second quantity is the double difference \begin{equation} {\rm ES}2 = (M_\Omega - M_\Delta) - 3 (M_{\Xi^*} - M_{\Sigma^*}) \,, \end{equation} which is one measure of the breaking of the equal spacing rule for decuplets. This is a good window on artifacts due to quenching because its expansion begins at $O(M_\pi^5)$ in QCD, but contains terms proportional to $\delta M_\pi^2 \log M_\pi$ in QQCD \cite{LS}. The LANL group finds that ${\rm ES}2$ differs from zero by 2-$\sigma$ \cite{ourspect96}, and I find that the data can be fit with $\delta\approx 0.1$ \cite{sharpebary}. In my view, the preponderance of the evidence suggests a value of $\delta$ in the range $0.1-0.2$. All extractions are complicated by the fact that the effects proportional to $\delta$ are small with present quark masses. To avoid them, one should use quark masses above $m_s/4-m_s/3$. This is true not only for the light quark quantities discussed above, but also for heavy-light quantities such as $f_B$. This, too, is predicted to be singular as the light quark mass vanishes \cite{zhang}. \section{QUENCHING ERRORS} \label{sec:querr} I close the first part of the talk by listing, in Table \ref{tab:querr}, a sampling of estimates of quenching errors, defined by \begin{equation} {\rm Error}({\rm Qty}) = { [ {\rm Qty}({\rm QCD})- {\rm Qty}({\rm QQCD})] \over{\rm Qty}({\rm QCD})} \,. \end{equation} I make the estimates by taking the numerator to be the difference between the pion loop contributions in the full and quenched chiral expansions. To obtain numerical values I set $\Lambda=m_\rho$ ($\Lambda$ is the scale occurring in chiral logs), use $f=f^Q=f_K$, and assume $\delta=0.1$ and $\alpha_\Phi=0$. For the estimates of heavy-light quantities I set $g'=0$, where $g'$ is an $\eta'$-$B$-$B$ coupling defined in Ref. \cite{zhang}. These estimates assume that the extrapolation to the light quark mass is done linearly from $m_q\approx m_s/2$. For example, $f_{B_d}$ in QQCD is {\em not} the quenched value with the physical $d$-quark mass (which would contain a large artifact proportional to $\delta$), but rather the value obtained by linear extrapolation from $m_s/2$, where the $\delta$ terms are much smaller. This is an attempt to mimic what is actually done in numerical simulations. \begin{table}[tb] \caption{Estimates of quenching errors.} \label{tab:querr} \begin{tabular}{ccl} \hline Qty. &Ref. & Error \\ \hline $f_\pi/M_\rho$ &\cite{gassleut,sharpechbk} & $\,\sim 0.1$ \\ $f_K/f_\pi-1$ & \cite{BGI} & $\ 0.4$ \\ $f_{B_s}$ & \cite{zhang} & $\ 0.2$\\ $f_{B_s}/f_{B_d}$ & \cite{zhang} & $\ 0.16$ \\ $B_{B_s}/B_{B_d}$ & \cite{zhang} & $\,-0.04$ \\ $B_K$ ($m_d=m_s$) & \cite{sharpechbk} & $\ 0$ \\ $B_K$ ($m_d\ne m_s$) & \cite{sharpetasi} & $\ 0.05$ \\ $M_\Xi-M_\Sigma$& \cite{sharpebary} & $\ 0.4$ \\ $M_\Sigma-M_N$ & \cite{sharpebary} & $\ 0.3$ \\ $M_\Omega-M_\Delta$& \cite{sharpebary} & $\ 0.3$ \\ \hline \end{tabular} \vspace{-0.2truein} \end{table} For the first two estimates, I have used the facts that, in QCD, \cite{gassleut} \begin{eqnarray} f_\pi &\approx& f\, [1 - 0.5 L(M_K)] \,,\\ f_K/f_\pi &\approx& 1 - 0.25 L(M_K) - 0.375 L(M_\eta) \,, \end{eqnarray} (where $L(M) = (M/4\pi f)^2 \log(M^2/\Lambda^2)$, and $f_\pi= 93\,$MeV), while in QQCD \cite{BGI,sharpechbk} \begin{eqnarray} f_\pi &\approx& f^Q \,,\\ {f_K\over f_\pi} &\approx& 1 + {\delta \over 2} \left[ {M_K^2 \over M_{ss}^2 - M_\pi^2} \log {M_{ss}^2 \over M_\pi^2} - 1 \right] \,. \end{eqnarray} I have not included the difference of pion loop contributions to $M_\rho$, since the loop has not been evaluated in QChPT, and a model calculation suggests that the difference is small \cite{cohenleinweber}. Details of the remaining estimates can be found in the references. Let me stress that these are estimates and not calculations. What they give is a sense of the effect of quenching on the contributions of ``pion'' clouds surrounding hadrons---these clouds are very different in QQCD and QCD! But this difference in clouds could be cancelled numerically by differences in the analytic terms in the chiral expansion. As discussed in Ref. \cite{zhang}, a more conservative view is thus to treat the estimates as rough upper bounds on the quenching error. Those involving ratios (e.g. $f_K/f_\pi$) are probably more reliable since some of the analytic terms do not contribute. One can also form double ratios for which the error estimates are yet more reliable (e.g. $R_{BG}$ and ES2 from the previous section; see also Ref. \cite{zhang}), but these quantities are of less phenomenological interest. My aim in making these estimates is to obtain a sense of which quenched quantities are likely to be more reliable and which less, and to get an sense of the possible size of quenching errors. My conclusion is that the errors could be significant in a number of quantities, including those involving heavy-light mesons. One might have hoped that the ratio $f_{B_s}/f_{B_d}$ would have small quenching errors, but the chiral loops indicate otherwise. For some other quantities, such as $B_{B_s}/B_{B_d}$ and $B_K$, the quenching errors are likely to be smaller. If these estimates work, then it will be worthwhile extending them to other matrix elements of phenomenological interest, e.g. $K\to\pi\pi$ amplitudes. Then, when numerical results in QQCD are obtained, we have at least a rough estimate of the quenching error in hand. Do the estimates work? As we will see below, those for $f_\pi/m_\rho$, $f_K/f_\pi$ and $B_K$ are consistent with the numerical results obtained to date. \section{RESULTS FOR DECAY CONSTANTS} \label{sec:decayc} For the remainder of the talk I will don the hat of a reviewer, and discuss the status of results for weak matrix elements. All results will be quenched, unless otherwise noted. I begin with $f_\pi/M_\rho$, the results for which are shown in Figs. \ref{fig:fpi_mrhoW} (Wilson fermions) and \ref{fig:fpi_mrhoCL} (SW fermions, with tadpole improved $c_{SW}$). The normalization here is $f_\pi^{\rm expt}=0.131\,$MeV, whereas I use $93\,$MeV elsewhere in this talk. \begin{figure}[tb] \vspace{-0.6truein} \centerline{\psfig{file=fig7.ps,height=2.5truein}} \vspace{-0.6truein} \caption{$f_\pi/M_\rho$ with quenched Wilson fermions.} \vspace{-0.2truein} \label{fig:fpi_mrhoW} \end{figure} Consider the Wilson data first. One expects a linear dependence on $a$, and the two lines are linear extrapolations taken from Ref. \cite{guptafpi}. The solid line is a fit to all the data, while the dashed curve excludes the point with largest $a$ (which might lie outside the linear region). It appears that the quenched result is lower than experiment, but there is a $5-10\%$ uncertainty. Improving the fermion action (Fig. \ref{fig:fpi_mrhoCL}) doesn't help much because of uncertainties in the normalization of the axial current. For the FNAL data, the upper (lower) points correspond to using $\alpha_s(\pi/a)$ ($\alpha_s(1/a)$) in the matching factor. The two sets of UKQCD95 points correspond to different normalization schemes. Again the results appear to extrapolate to a point below experiment. \begin{figure}[tb] \vspace{-0.1truein} \centerline{\psfig{file=fig8.ps,height=3.truein}} \vspace{-0.6truein} \caption{$f_\pi/M_\rho$ with quenched SW fermions.} \vspace{-0.2truein} \label{fig:fpi_mrhoCL} \end{figure} It is disappointing that we have not done better with such a basic quantity. We need to reduce both statistical errors and normalization uncertainty. The latter may require non-perturbative methods, or the use of staggered fermions (where $Z_A=1$). Note that chiral loops estimate that the quenched result will undershoot by 12\%, and this appears correct in sign, and not far off in magnitude. Results for $(f_K-f_\pi)/f_\pi$ are shown in Fig. \ref{fig:fk_fpi}. This ratio measures the mass dependence of decay constants. Chiral loops suggest a 40\% underestimate in QQCD. The line is a fit to all the Wilson data (including the largest $a$'s), and indeed gives a result about half of the experimental value. The new UKQCD results, using tadpole improved SW fermions, are, by contrast, rising towards the experimental value. It will take a substantial reduction in statistical errors to sort this out. \begin{figure}[tb] \vspace{-0.1truein} \centerline{\psfig{file=fig9.ps,height=3.0truein}} \vspace{-0.6truein} \caption{$(f_K-f_\pi)/f_\pi$ in quenched QCD.} \vspace{-0.2truein} \label{fig:fk_fpi} \end{figure} \section{STATUS OF $B_K$: STAGGERED} \label{sec:bks} $B_K$ is defined by \begin{equation} B_K = {\langle \bar K| \bar s \gamma_\mu^L d\, \bar s\gamma_\mu^L d | K \rangle \over (8/3) \langle \bar K| \bar s \gamma_\mu^L d|0 \rangle \langle 0 |\bar s\gamma_\mu^L d | K \rangle } \,. \label{eq:bkdef} \end{equation} It is a scale dependent quantity, and I will quote results in the NDR (naive dimensional regularization) scheme at 2 GeV. It can be calculated with very small statistical errors, and has turned out to be a fount of knowledge about systematic errors. This is true for both staggered and Wilson fermions, though for different reasons. There has been considerable progress with both types of fermions in the last year. I begin with staggered fermions, which hold the advantage for $B_K$ as they have a remnant chiral symmetry. Back in 1989, I thought we knew what the quenched answer was, based on calculations at $\beta=6$ on $16^3$ and $24^3$ lattices: $B_K = 0.70(2)$ \cite{sharpe89,ourbkprl}. I also argued that quenching errors were likely small (see Table \ref{tab:querr}). I was wrong on the former, though maybe not on the latter. By 1993, Gupta, Kilcup and I had found that $B_K$ had a considerable $a$ dependence \cite{sharpe93}. Applying Symanzik's improvement program, I argued that the discretization errors in $B_K$ should be $O(a^2)$, and not $O(a)$. Based on this, we extrapolated our data quadratically, and quoted $B_K(NDR,2{\rm GeV}) = 0.616(20)(27)$ for the quenched result. Our data alone, however, was not good enough to distinguish linear and quadratic dependences. Last year, JLQCD presented results from a more extensive study (using $\beta=5.85$, $5.93$, $6$ and $6.2$) \cite{jlqcdbk95}. Their data strongly favored a linear dependence on $a$. If correct, this would lead to a value of $B_K$ close to $0.5$. The only hope for someone convinced of an $a^2$ dependence was competition between a number of terms. Faced with this contradiction between numerical data and theory, JLQCD have done further work on both fronts \cite{aoki96}. They have added two additional lattice spacings, $\beta=5.7$ and $6.4$, thus increasing the lever arm. They have also carried out finite volume studies at $\beta=6$ and $6.4$, finding only a small effect. Their data are shown in Fig. \ref{fig:jlqcdbk}. ``Invariant'' and ``Landau'' refer to two possible discretizations of the operators---the staggered fermion operators are spread out over a $2^4$ hypercube, and one can either make them gauge invariant by including gauge links, or by fixing to Landau gauge and omitting the links. The solid (dashed) lines show quadratic (linear) fits to the first five points. The $\chi^2/{\rm d.o.f.}$ are \begin{center} \begin{tabular}{ccc} \hline Fit & Invariant & Landau \\ $a$ & 0.86 & 0.67 \\ $a^2$ & 1.80 & 2.21 \\ \hline \end{tabular} \end{center} thus favoring the linear fit, but by a much smaller difference than last year. If one uses only the first four points then linear and quadratic fits are equally good. What has changed since last year is that the new point at $\beta=6.4$ lies above the straight line passing through the next four points. \begin{figure}[tb] \vspace{-0.1truein} \centerline{\psfig{file=fig10.ps,height=3.0truein}} \vspace{-0.6truein} \caption{JLQCD results for staggered $B_K$.} \vspace{-0.2truein} \label{fig:jlqcdbk} \end{figure} JLQCD have also checked the theoretical argument using a simpler method of operator enumeration\cite{aoki96,ishizuka96}.\footnote{ A similar method has also been introduced by Luo \cite{luo96}.} The conclusion is that there cannot be $O(a)$ corrections to $B_K$, because there are no operators available with which one could remove these corrections. Thus JLQCD use quadratic extrapolation and quote (for degenerate quarks) \begin{equation} B_K({\rm NDR}, 2\,{\rm GeV}) = 0.5977 \pm 0.0064 \pm 0.0166 \,, \label{eq:jlqcdbk} \end{equation} where the first error is statistical, the second due to truncation of perturbation theory. This new result agrees with that from 1993 (indeed, the results are consistent at each $\beta$), but has much smaller errors. To give an indication of how far things have come, compare our 1993 result at $\beta=6$ with Landau-gauge operators, $0.723(87)$ \cite{sharpe93}, to the corresponding JLQCD result $0.714(12)$. The perturbative error in $B_K$ arises from truncating the matching of lattice and continuum operators to one-loop order. The use of two different lattice operators allows one to estimate this error without resort to guesswork about the higher order terms in the perturbative expansion. The difference between the results from the two operators is of $O[\alpha(2\,{\rm GeV})^2]$, and thus should remain finite in the continuum limit. This is what is observed in Fig. \ref{fig:jlqcdbk}. I will take Eq. \ref{eq:jlqcdbk} as the best estimate of $B_K$ in QQCD. The errors are so much smaller than those in previous staggered results and in the results with Wilson fermions discussed below, that the global average is not significantly different from the JLQCD number alone. The saga is not quite over, however, since one should confirm the $a^2$ dependence by running at even smaller lattice spacings. JLQCD intend to run at $\beta=6.6$. If the present extrapolation holds up, then it shows how one must beware of keeping only a single term when extrapolating in $a$. \subsection{Unquenching $B_K$} To obtain a result for QCD proper, two steps remain: the inclusion of dynamical quarks, and the use of $m_s\ne m_d$. The OSU group has made important progress on the first step \cite{osubk96}. Previous studies (summarized in Ref. \cite{soni95}) found that $B_K$ was reduced slightly by sea quarks, although the effect was not statistically significant. The OSU study, by contrast, finds a statistically significant increase in $B_K$ \begin{equation} { B_K({\rm NDR,2\,GeV},N_f=3) \over B_K({\rm NDR,2\,GeV},N_f=0)} = 1.05 \pm 0.02 \,. \label{eq:bkquerr} \end{equation} They have improved upon previous work by reducing statistical errors, and by choosing their point at $N_f=0$ ($\beta=6.05$) to better match the lattice spacing at $N_f=2$ ($\beta=5.7$, $m_qa=0.01$) and $4$ ($\beta=5.4$, $m_qa=0.01$). There are systematic errors in this result which have yet to be estimated. First, the dynamical lattices are chosen to have $m_q^{\rm sea}=m_q^{\rm val} = m_s^{\rm phys}/2$, and so they are truely unquenched simulations. But $m_s^{\rm phys}$ is determined by extrapolating in the valence quark mass alone, and is thus a partially quenched result. This introduces an uncertainty in $m_s$ which feeds into the estimate of the $N_f$ dependence of $B_K$. Similarly, $a$ is determined by a partially quenched extrapolation, resulting in an uncertainty in the matching factors between lattice and continuum operators. But probably the most important error comes from the possibility of significant $a$ dependence in the ratio in Eq. \ref{eq:bkquerr}. The result quoted is for $a^{-1}=2\,$GeV, at which $a$ the discretization error in the quenched $B_K$ is 15\%. It is not inconceivable that, say, $B_K$ in QCD has very little dependence on $a$, in which case the ratio would increase to $\sim 1.2$ in the continuum limit. Clearly it is very important to repeat the comparison at a different lattice spacing. Despite these uncertainties, I will take the OSU result and error as the best estimate of the effect of quenching at $a=0$. I am being less conservative than I might be because a small quenching error in $B_K$ is consistent with the expectations of QChPT. A more conservative estimate for the ratio would be $1.05\pm0.15$. \subsection{$B_K$ for non-degenerate quarks} \label{subsec:bknondegen} What remains is to extrapolate from $m_s=m_d\approx m_s^{\rm phys}/2$ to $m_s=m_s^{\rm phys}$ and $m_d=m_d^{\rm phys}$. This appears difficult because it requires dynamical quarks with very small masses. This may not be necessary, however, if one uses ChPT to guide the extrapolation \cite{sharpetasi}. The point is that the chiral expansion in QCD is \cite{bijnens,sharpechbk} \begin{equation} {B_K\over B} = 1 - \left(3+{\epsilon^2 \over 3}\right) y\ln{y} + b y + c y \epsilon^2 \,, \label{eq:bkchqcd} \end{equation} where \begin{equation} \epsilon=(m_s-m_d)/(m_s+m_d)\,,\ y = M_K^2/(4 \pi f)^2 , \end{equation} and $B$, $b$ and $c$ are unknown constants. At this order $f$ can be equally well taken to be $f_\pi$ or $f_K$. Equation \ref{eq:bkchqcd} is an expansion in $y$, but is valid for all $\epsilon$. The idea is to determine $c$ by working at small $\epsilon$, and then use the formula to extrapolate to $\epsilon=1$. This ignores corrections of $O(y^2)$, and so the errors in the extrapolation are likely to be $\sim 25\%$. Notice that $m_u$ does not enter into Eq. \ref{eq:bkchqcd}. Thus one can get away with a simulation using only two different dynamical quark masses, e.g. setting $m_u=m_d < m_s^{\rm phys}/2$, while holding $m_s +m_d = m_s^{\rm phys}$. To date, no such calculation has been done. To make an estimate I use the chiral log alone, i.e. set $c=0$, yielding \begin{equation} B_K({\rm non-degen}) = (1.04-1.08) B_K({\rm degen}) \,. \end{equation} The range comes from using $f=f_\pi$ and $f_K$, and varying the scale in the logarithm from $m_\rho-1\,$GeV. Since the chiral log comes mainly from kaon and $\eta$ loops \cite{sharpechbk}, I prefer $f=f_K$, which leads to $1.04-1.05$ for the ratio. To be conservative I take $1.05\pm0.05$, and assume that the generous error is large enough to include also the error in the estimate of the effect of unquenching. This leads to a final estimate of \begin{equation} B_K({\rm NDR},{\rm 2\,GeV,QCD}) = 0.66 \pm 0.02 \pm 0.03 \,, \label{eq:finalqcdbk} \end{equation} where the first error is that in the quenched value, the second that in the estimate of unquenching and using non-degenerate quarks. Taking the more conservative estimate of the unquenching error (15\%), and adding it in quadrature with the (5\%) estimate of the error in accounting for non-degenerate quarks, increases the second error in Eq. \ref{eq:finalqcdbk} to $0.11$. It is customary to quote a result for the renormalization group invariant quantity \[ {\widehat{B}_K \over B_K(\mu)} = \alpha_s(\mu)^{-\gamma_0 \over2\beta_0} \left(1 + {\alpha_s(\mu) \over 4 \pi} \left[{\beta_1 \gamma_0 -\beta_0\gamma_1 \over 2 \beta_0^2} \right] \right) \] in the notation of Ref. \cite{crisafulli}. Using $\alpha_s(2\,{\rm GeV})=0.3$ and $N_f=3$, I find $\widehat{B}_K=0.90(3)(4)$, with the last error increasing to $0.14$ with the more conservative error. This differs from the result I quoted in Ref. \cite{sharpe93}, because I am here using the 2-loop formula and a continuum choice of $\alpha_s$. \subsection{Chiral behavior of $B_K$} \label{subsec:bkchfit} Since $B_K$ can be calculated very accurately, it provides a potential testing ground for (partially) quenched ChPT. This year, for the first time, such tests have been undertaken, with results from OSU \cite{osubk96}, JLQCD \cite{aoki96}, and Lee and Klomfass \cite{lee96}. I note only some highlights. It turns out that, for $\epsilon=0$, Eq. \ref{eq:bkchqcd} is valid for all $N_f$ \cite{sharpechbk}. This is why my estimate of the quenching error for $B_K$ with degenerate quarks in Table \ref{tab:querr} is zero. Thus the first test of (P)QChPT is to see whether the $-3 y\ln y$ term is present. The OSU group has the most extensive data as a function of $y$, and indeed observe curvature of the expected sign and magnitude for $N_f=0,2,4$. JLQCD also finds reasonable fits to the chiral form, as long as they allow a substantial dependence of $f$ on lattice spacing. They also study other $B$ parameters, with similar conclusions. Not everything works. JLQCD finds that the volume dependence predicted by the chiral log \cite{sharpechbk} is too small to fit their data. Fitting to the expected form for $\epsilon\ne0$ in QQCD, they find $\delta=-0.3(3)$, i.e. of the opposite sign to the other determinations discussed in Sec. \ref{sec:chevidence}. Lee and Klomfass have studied the $\epsilon$ dependence with $N_f=2$ (for which there is as yet no PQChPT prediction). It will be interesting to see how things evolve. My only comment is that one may need to include $O(y^2)$ terms in the chiral fits. \section{STATUS OF $B_K$: WILSON} \label{sec:bkw} There has also been considerable progress in the last year in the calculation of $B_K$ using Wilson and SW fermions. The challenge here is to account for the effects of the explicit chiral symmetry breaking in the fermion action. Success with $B_K$ would give one confidence to attempt more complicated calculations. The operator of interest, \begin{equation} {\cal O}_{V+A} = \bar s \gamma_\mu d\, \bar s \gamma_\mu d + \bar s \gamma_\mu \gamma_5 d\, \bar s \gamma_\mu \gamma_5 d \,, \end{equation} can ``mix'' with four other dimension 6 operators \begin{equation} {\cal O}_{V+A}^{\rm cont} = Z_{V+A} \left( {\cal O}_{V+A} + \sum_{i=1}^{4} z_i {\cal O}_i \right) + O(a) \end{equation} where the ${\cal O}$ on the r.h.s. are lattice operators. The ${\cal O}_i$ are listed in Refs. \cite{kuramashi96,talevi96}. The meaning of this equation is that, for the appropriate choices of $Z_{V+A}$ and the $z_i$, the lattice and continuum operators will have the same matrix elements, up to corrections of $O(a)$. In particular, while the matrix elements of a general four fermion operator has the chiral expansion \begin{eqnarray} \lefteqn{\langle \bar K | {\cal O} | K \rangle = \alpha + \beta M_K^2 + \delta_1 M_K^4 +} \\ && p_{\bar K}\cdot p_{K} (\gamma + \delta_2 M_K^2 + \delta_3 p_{\bar K}\cdot p_{K} ) +\dots \,, \end{eqnarray} chiral symmetry implies that $\alpha=\beta=\delta_1=0$ for the particular operator ${\cal O}={\cal O}_{V+A}^{\rm cont}$. Thus, one can test that the $z_i$ are correct by checking that the first three terms are absent.\footnote{I have ignored chiral logarithms, which will complicate the analysis, but can probably be ignored given present errors and ranges of $M_K$.} Note that the $z_i$ must be known quite accurately because the terms we are removing are higher order in the chiral expansion than the terms we are keeping. Five methods have been used to determine the $z_i$ and $Z_{V+A}$. \\ (1) One-loop perturbation theory. This fails to give the correct chiral behavior, even when tadpole improved. \\ (2) Use (1) plus enforce chiral behavior by adjusting subsets of the $z_i$ by hand \cite{bernardsoni89}. Different subsets give differing results, introducing an additional systematic error. Results for a variety of $a$ were presented by Soni last year \cite{soni95}.\\ (3) Use (1) and discard the $\alpha$, $\beta$ and $\gamma$ terms, determined by doing the calculation at a variety of momenta. New results using this method come from the LANL group \cite{gupta96}. Since the $z_i$ are incorrect, there is, however, an error of $O(g^4)$ in $B_K$. \\ (4) Non-perturbative matching by imposing continuum normalization conditions on Landau-gauge quark matrix elements. This approach has been pioneered by the Rome group, and is reviewed here by Rossi \cite{rossi96}. The original calculation omitted one operator \cite{donini}, but has now been corrected \cite{talevi96}. \\ (5) Determine the $z_i$ non-perturbatively by imposing chiral ward identities on quark matrix elements. Determine $Z_{V+A}$ as in (4). This method has been introduced by JLQCD \cite{kuramashi96}. The methods of choice are clearly (4) and (5), as long as they can determine the $z_i$ accurately enough. In fact, both methods work well: the errors in the non-perturbative results are much smaller than their difference from the one-loop perturbative values. And both methods find that the matrix element of ${\cal O}_{V+A}^{\rm cont}$ has the correct chiral behavior, within statistical errors. What remains to be studied is the uncertainty introduced by the fact that there are Gribov copies in Landau gauge. Prior experience suggests that this will be a small effect. It is not yet clear which, if either, of methods (4) and (5) is preferable for determining the $z_i$. As stressed in Ref. \cite{talevi96} the $z_i$ are unique, up to corrections of $O(a)$. In this sense, both methods must give the same results. But they are quite different in detail, and it may be that the errors are smaller with one method or the other. It will be interesting to see a comprehensive comparison between them and also with perturbation theory. In Fig. \ref{fig:bkw} I collect the results for $B_K$. All calculations use $m_s=m_d$ and the quenched approximation. The fact that most of the results agree is a significant success, given the variety of methods employed. It is hard to judge which method gives the smallest errors, because each group uses different ensembles and lattice sizes, and estimates systematic errors differently. The errors are larger than with staggered fermions mostly because of the errors in the $z_i$. \begin{figure}[tb] \vspace{-0.1truein} \centerline{\psfig{file=fig11.ps,height=3.0truein}} \vspace{-0.6truein} \caption{Quenched $B_K$ with Wilson fermions.} \vspace{-0.2truein} \label{fig:bkw} \end{figure} Extrapolating to $a=0$ using the data in Fig. \ref{fig:bkw} would give a result with a large uncertainty. Fortunately, JLQCD has found a more accurate approach. Instead of $B_K$, they consider the ratio of the matrix element of ${\cal O}_{V+A}^{\rm cont}$ to its vacuum saturation approximant. The latter differs from the denominator of $B_K$ (Eq. \ref{eq:bkdef}) at finite lattice spacing. The advantage of this choice is that the $z_i$ appear in both the numerator and denominator, leading to smaller statistical errors. The disadvantage is that the new ratio has the wrong chiral behavior at finite $a$. It turns out that there is an overall gain, and from their calculations at $\beta=5.9$, $6.1$ and $6.3$ they find $B_K({\rm NDR,2\,GeV})=0.63(8)$. This is the result shown at $a=0$ in Fig. \ref{fig:bkw}. It agrees with the staggered result, although it has much larger errors. Nevertheless, it is an important consistency check, and is close to ruling out the use of a linear extrapolation in $a$ with staggered fermions. \section{OTHER MATRIX ELEMENTS} \label{sec:otherme} The LANL group \cite{gupta96} has quenched results (at $\beta=6$) for the matrix elements which determine the dominant part of the electromagnetic penguin contribution to $\epsilon'/\epsilon$ \begin{eqnarray} B_7^{I=3/2} &=& 0.58 \pm 0.02 {\rm (stat)} {+0.07 \atop -0.03} \,, \\ B_8^{I=3/2} &=& 0.81 \pm 0.03 {\rm (stat)} {+0.03 \atop -0.02} \,. \end{eqnarray} These are in the NDR scheme at 2 GeV. The second error is from the truncation of perturbation matching factors. These numbers lie at or below the lower end of the range used by phenomenologists. The LANL group also finds $B_D=0.78(1)$. There are also new results for $f_\rho$ and $f_\phi$ \cite{guptafpi,yoshie96}, for the pion polarizability \cite{wilcox96}, and for strange quark contributions to magnetic moments \cite{dong96}. \section{FUTURE DIRECTIONS} \label{sec:future} This year has seen the first detailed tests of the predicted chiral behavior of quenched quantities. Further work along these lines will help us make better extrapolations, and improve our understanding of quenching errors. It is also a warm-up exercise for the use of chiral perturbation theory in unquenched theories. I have outlined one such application in Sec. \ref{subsec:bknondegen}. I expect the technique to be of wide utility given the difficulty in simulating light dynamical fermions. As for matrix elements, there has been substantial progress on $B_K$. It appears that we finally know the quenched result, thanks largely to the efforts of JLQCD. At the same time, it is disturbing that the complicated $a$ dependence has made it so difficult to remove the last 20\% of the errors. One wonders whether similar complications lie lurking beneath the relatively large errors in other matrix elements. The improved results for $B_K$ with Wilson fermions show that non-perturbative normalization of operators is viable. My hope is that we can now return to an issue set aside in 1989: the calculation of $K\to\pi\pi$ amplitudes. The main stumbling block is the need to subtract lower-dimension operators. A method exists for staggered fermions, but the errors have so far swamped the signal. With Wilson fermions, one needs a non-perturbative method, and the hope is that using quark matrix elements in Landau gauge will do the job \cite{talevi96}. Work is underway with both types of fermion. Given the success of the Schr\"odinger functional method at calculating current renormalizations \cite{wittig96}, it should be tried also for four fermion operators. Back in 1989, I also described preliminary work on non-leptonic $D$ decays, e.g. $D\to K\pi$. Almost no progress has been made since then, largely because we have been lacking a good model of the decay amplitude for Euclidean momenta. A recent proposal by Ciuchini {\em et al.} may fill this gap \cite{ciuchini}. Enormous computational resources have been used to calculate matrix elements in (P)QQCD. To proceed to QCD at anything other than a snail's pace may well require the use of improved actions. Indeed, the large discretization errors in quenched staggered $B_K$ already cry out for improvement. The fact that we know $B_K$ very accurately will provide an excellent benchmark for such calculations. Working at smaller values of the cut-off, $1/a$, alleviates some problems while making others worse. Subtraction of lower dimension operators becomes simpler, but the evaluation of mixing with operators of the same dimension becomes more difficult. It will be very interesting to see how things develop. \section*{Acknowledgements} I am grateful to Peter Lepage for helpful conversations and comments on the manuscript. \def\PRL#1#2#3{{Phys. Rev. Lett.} {\bf #1}, #3 (#2)} \def\PRD#1#2#3{{Phys. Rev.} {\bf D#1}, #3 (#2)} \def\PLB#1#2#3{{Phys. Lett.} {\bf #1B} (#2) #3} \def\NPB#1#2#3{{Nucl. Phys.} {\bf B#1} (#2) #3} \def\NPBPS#1#2#3{{Nucl. Phys.} {\bf B({Proc.Suppl.}) {#1}} (#2) #3} \def{\em et al.}{{\em et al.}}
proofpile-arXiv_065-421
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} In papers \cite{1} was shown that the theory of integrable systems (under the assumption of commutativety of all involved functions) can be reformulated in a form, where the key role plays the group of integrable mappings and its theory of representation. It arose the question, what will happen with this construction, if we will consider equations of motion for Heisenberg operators, or in other words, when unknown functions of integrable systems changed on non--commutative variables? The goal of the present paper is to give the partial answer to this question. Each quantum system with the same success can be described in many different (in form) representations. The most known and used are Schr\"odinger and Heisenberg pictures. The first deal with the wave functions (the state vectors in a Hilbert space), the second with the non--commutative Heisenberg operators and equations of motion under appropriate initial conditions (commutation relations at the fixed moment of time). In this paper we will show how the group of integrable mappings conception must be changed to include the non--commutative variables case. The equations of evolution type (after some modifications connected with the order of a multipliers) remain invariant with respect to the corresponding quantum discrete transformations without any assumption about commutation rules for unknown functions (operators). Partially they can be $s\times s$ matrix functions or some operators acting in the arbitrary representation space.The equations of motion for quantum Heisenberg operators are containing within this construction. We use the discrete transformations method as the most adequate to solution of such kind problems \cite{1}. We restrict ourselves by some concrete examples of integrable mappings for non--commutative objects, in usual and supersymmetrical two--dimensional spaces and by corresponding hierarchies of $(1+2)$--dimensional integrable systems. We now have no idea how to enumerate all possible integrable mappings for non--commutative objects. In this connection we can only add that in commutative case this problem yet is very far {}From its final solution. \section{Non--Commutative Darboux--Toda Substitution in Two--Dimensional space} \subsection{Definitions} Let us denote $u,v$ the pair of operators defined in some representation space and depending on $x,y$ coordinates of two--dimension space. We assume, that partial derivatives up to some sufficient large order and inverse operators $u^{-1}, v^{-1}$ are exist. Only associativety is assumed for multiplication, nowhere we will assume any commutation relations. We will consider the following mapping: \begin{equation} \begin{array}{cc} \lef{u}{}= v^{-1}&\lef{v}{}= [vu-(v_x v^{-1})_y]v\equiv v[uv- (v^{-1} v_y)_x], \end{array} \label{10} \end{equation} {\emergencystretch=5pt where $\lef{u}{}; \lef{v}{}$ denotes final, transformed operators. In the case, when $u,v$ are some $s\times s$ matrices (\ref{10}) was considered in \cite{10}. In classical case, when $u,v$ are usual commutative functions (\ref{10}) is the well-known Darboux--Toda substitution. Substitution (\ref{10}) is invertible, i.e. the initial operators can be expressed in the terms of final ones. Denoting $\rig{u}{}; \rig{v}{}$ the result of inverse transformation we have: \begin{equation} \begin{array}{cc} \rig{v}{}= u^{-1}&\rig{u}{}= [uv-(u_y u^{-1})_x]u\equiv u[vu- (u^{-1}u_y)_x] \end{array} \label{11} \end{equation} Operator $ f(u,v)$ after application of the s--times direct transformation we will denote $ \lef{f}{s}\equiv f(\lef{u}{s},\lef{v}{s})$ and after s--times inverse transformation as $\rig{f}{s}~\equiv~f~(\rig{u}{s}~,~\rig{v}{s}~)$, with the agreement $\lef{f}{-m}\equiv\rig{f}{m}$, $m\ge0$. } \noindent If \begin{equation} \begin{array}{c} u_t=F_1(u,v,u_x,v_x,u_y,v_y,...)\\ v_t=F_2(u,v,u_x,v_x,u_y,v_y,...) \end{array}\label{611} \end{equation} is given evolution type system then the condition of its invariance with respect to the transformation (\ref{10}) (it means that in terms of $\lef{u}{},\lef{v}{}$ equations (\ref{611}) will be exactly the same as they are in terms of $u,v$) can be derived by differentiation of (\ref{10}) by some parameter and has the following form: \begin{equation} \begin{array}{rl} \lef{F}{}_1= \lef{u}{}_t=&-v^{-1}v_t v^{-1}=-v^{-1}F_2 v^{-1}\\ \lef{F}{}_2= \lef{v}{}_t=&([vu-(v_x v^{-1})_y]v)_t=[F_2 u+v F_1-(F_{2x}v^{-1})_y+\\ &+(v_xv^{-1}F_2v^{-1})_y]v+[vu-(v_xv^{-1})_y]F_2 \end{array} \label{13} \end{equation} This is the functional symmetry equation for substitution (\ref{10}). Unknown operators here are $F_1, F_2$. If some operators $F_1(u,v), F_2(u,v)$ are solution of (\ref{13}) then the corresponding system (\ref{611}) will be invariant with respect to (\ref{10}). (\ref{13}) is a linear system, i.e. if ${F_1}', {F_2}'$ and ${F_1}'', {F_2}''$ are solutions then $F_1= a{F_1}'+b{F_1}'', F_2=a{F_2}'+b{F_2}''$, where $a, b$ are arbitrary numerical parameters, is also solution. Every symmetry equation possesses trivial solution $F_1=au_x+bu_y, F_2=av_x+bv_y$. Substitution is called integrable if its symmetry equation have at least one non--trivial solution. \subsection{Solution of the Symmetry Equation} The method we will use here to find solutions of (\ref{13}) is analogues to the method we used in \cite{14} in the case of commutative functions. But it is not exactly the same as many transformations can not be done because of non--commutativety of variables under consideration. First of all let us take $F_2=\alpha_0 v$, $F_1=u\beta_0$. We obtain: \begin{equation} \begin{array}{l} \beta_0=-\rig{\alpha}{}_0\\ {\alpha_0}_{xy}=(\alpha_0-\lef{\alpha}{}_0)\lef{T_0}{\:}+ T_0(\alpha_0-\rig{\alpha_0}{\!})+\theta {\alpha_0}_y, -{\alpha_0}_y\theta \label{2} \end{array} \end{equation} where $T_0=vu$, $\theta=v_x v^{-1}$. This system possesses obvious partial solution ${\alpha_0}^{(0)}=-{\beta_0}^{(0)}=1$, which gives the first term of hierarchy: $F_1=-u, F_2=v$. Two following equations for $ T_0 $ and $ \theta $, which are the direct corollary of (\ref{10}), are important for father calculations: \begin{equation} {T_0}_x=\theta T_0-T_0 \rig{\theta}{}\qquad \theta_y=T_0-\lef{T_0}{\:} \label{3} \end{equation} In fact, (\ref{3}) is the substitution (\ref{10}) rewritten in terms of $T_0$ and $\theta$. Now let us take ${\alpha_0}_y=\alpha_1\lef{T_0}{\:}+T_0\beta_1$. One can treat this expression as analog to the decomposition of some vector by basic vectors. {}From (\ref{2}), expressing $\lef{T_0}{\:}_x $ and $T_{0x}$ with the help of (\ref{3}) and equating to zero coefficients before $\lef{T}{}_0$ and after $T_0$ (which is some additional assumption), we have: \begin{equation} \begin{array}{l} {\alpha_1}_x=\alpha_0-\lef{\alpha}{}_0+\theta\alpha_1-\alpha_1\lef{\theta}{}\\ {\beta_1}_x=\alpha_0-\rig{\alpha_0}{\!}+\rig{\theta}{}\beta_1-\beta_1\theta \end{array}\label{1000} \end{equation} The second relation obviously can be rewritten as: $$ {\lef{\beta}{}_1}_x=-(\alpha_0-\lef{\alpha}{}_0)+\theta\lef{\beta}{}_1- \lef{\beta}{}_1\lef{\theta}{} $$ {}From what it follows that system (\ref{1000}) possesses partial solution of the form $\lef{\beta}{}_1=-\alpha_1$. After taking $y$--derivative of equation for $\alpha_1$, this partial solution gives the following system: \begin{equation} \begin{array}{l} \beta_1=-\rig{\alpha_1}{\!}\\ {\alpha_1}_{xy}=(\alpha_1-\lef{\alpha}{}_1)\lef{T_0}{2}+T_0(\alpha_1-\rig{\alpha_1}{\!})+ \theta{\alpha_1}_y -{\alpha_1}_y \lef{\theta}{}\\ \end{array} \label{4} \end{equation} This system is analogues to the (\ref{2}). It also has partial solution $\alpha_1=-\beta_1=1$, which leads to ${\alpha_0}^{(1)}=\int(\lef{T_0}{\:}-T_0)dy$, and gives the next solution of the symmetry equation (\ref{13}). Taking now ${\alpha_1}_y=\alpha_2\lef{T_0}{2}+T_0\beta_2$ we are able to continue by the same scheme. The system for $\alpha_2, \beta_2$ has the same structure as previous systems. Its partial solution $\alpha_2=-\beta_2$ allows to find the third term of hierarchy $$ {\alpha_0}^{(2)}=\int dy\,\Bigl[\Bigl(\int dy\,\Bigl(\lef{T_0}{2}-T_0\Bigr)\Bigr) \lef{T_0}{\:}-T_0\int dy\,\Bigl(\lef{T_0}{\:}-\rig{T_0}{\!}\Bigr)\Bigr] $$ By induction it can be proved that in general case equations for $\alpha_n, \beta_n $ have the form: \begin{equation} \begin{array}{l} \beta_n=-\rig{\alpha_n}{\!}\\ {\alpha_n}_{xy}=(\alpha_n-\lef{\alpha}{}_n)\lef{T_0}{(n+1)}+T_0(\alpha_n-\rig{\alpha_n}{\!}) +\theta{\alpha_n}_y-{\alpha_n}_{y}\lef{\theta}{n}\\ \end{array} \label{5} \end{equation} with partial solution $\alpha_n=-\beta_n=1$. After this expression for ${\alpha_0}^{(n)}$ can be reconstructed in the form of the sum of $2^n$ terms, which can be written in the following symbolical form: \begin{equation} \begin{array}{rcl} {\alpha_0}^{(n)}&=&(-1)^n \prod_{i=1}^n \left(1-L_i exp\left[id_i+\sum_{i=k+1}^n d_k\right]\right)\times \\&&\\ &&\times\int dy (T_0 \int dy (\rig{T_0}{\!}\int dy (...\int dy \rig{T_0}{(n-1)}...)))\label{20} \end{array} \end{equation} where $exp\, d_p$ means shifts by the unity the argument of p\--repeated integral. $$ \dots\int dy\lef{T_0}{p}\to\dots\int dy\lef{T_0}{(p+1)}\dots $$ and symbol $L_r$\---transposition of terms in the r\--th brackets $$ (A_1(\dots(A_r[\dots])\dots)) \to (A_1(\dots([\dots]A_r)\dots)) $$ with the following multiplication rules: $$ L_i exp[...]_1 L_j exp[...]_2=L_i L_j exp\left[ [...]_1+[...]_2\right] $$ Comparing (\ref{20}) with \cite{14} we see that here for non-commutativety we are forced to introduce the new operators $L_i$ which are discount the order of the multipliers. \subsection{Examples} \subsubsection*{n=0} $$ v_t=v\qquad u_t=-u $$ \subsubsection*{n=1} $$ v_t=v_x\qquad u_t=u_x $$ \subsubsection*{n=2} $$ v_t=v_{xx}-2\int (vu)_x dy\times v \quad u_t=-u_{xx}+2u\int (vu)_x dy $$ This is the Davey--Stewartson system, described in \cite{12} \subsubsection*{n=3} $$ v_t=v_{xxx}-3\int (vu)_x dy\times v_x-3\int (v_x u)_x dy\times v- $$ $$ -3\int \left[vu\int (vu)_x dy-\int (vu)_x dy\times vu\right]dy\times v $$ \vspace{1em} $$ u_t=-u_{xxx}-3u_x \int (vu)_x dy-3u\int (v_x u)_x dy- $$ $$ -3u\int \left[vu\int dy (vu)_x-\left( \int dy (vu)_x \right) vu\right]dy $$ In commutative case this is the Veselov--Novikov system. \section{Non--Commutative Darboux--Toda Transformation in Two--Dimensional Super Space} \subsection{Definitions} Here we will analyze the situation, when non-commutative operators under consideration in addition to usual space and time coordinates $x, y, t$ are depend upon Grassman variables $\theta_+, \theta_-$. We will consider the following mapping: \begin{equation} \begin{array}{cc} \lef{u}{}= v^{-1}&\lef{v}{}=-[D_-(D_+v\times v^{-1})+vu]v\equiv v[D_+(v^{-1}D_-v)-uv], \end{array} \label{31} \end{equation} where $$ D_+=\frac{\partial}{\partial\theta_+}+\theta_+\frac{\partial}{\partial x}\quad D_-=\frac{\partial}{\partial\theta_-}+\theta_-\frac{\partial}{\partial y}\quad D^2_+=\frac{\partial}{\partial x}\quad D^2_-=\frac{\partial}{\partial y} $$ Other notations are the same as in the previous section. Substitution (\ref{31}) is invertible. Inverse transformation has the form: \begin{equation} \begin{array}{cc} \rig{v}{}= u^{-1}&\rig{u}{}=-[D_+(D_-u\times u^{-1})+uv]u\equiv u[D_-(u^{-1}D_+u)-vu], \end{array} \label{33} \end{equation} The symmetry equation for (\ref{31}) is the following: \begin{equation} \begin{array}{rcl} \lef{F_1}{\:}&= & -v^{-1}F_2 v^{-1}\\ \lef{F_2}{\,}&=& F_2[D_+(v^{-1}D_-v)-uv]+v[D_+(-v^{-1}F_2v^{-1}D_-v)+\\&&+D_+(v^{-1}D_-F_2)-F_1v- uF_2] \end{array} \label{35} \end{equation} \subsection{Solution of the Symmetry Equation} Here we will get the hierarchy of solutions of the symmetry equation (\ref{35}). For this we will use the same general method as in the previous section. But there is an interesting and in some sense important difference. As we will see bellow, partial solutions of (\ref{35}) can be found only at even steps, when unknown operators are Bosonic--like variables, whereas at odd steps they are Fermionic--like. After substitution in (\ref{35}) $F_1=u\beta_0, F_2=\alpha_0 v$ we have: \begin{equation} \begin{array}{l} \beta_0=-\rig{\alpha_0}{\!}\\ D_+D_-\alpha_0=(\lef{\alpha}{}_0-\alpha_0)\lef{T_0}{\:}+ T_0(\alpha_0-\rig{\alpha_0}{\!})+\theta D_-\alpha_0+D_-\alpha_0\theta, \label{62} \end{array} \end{equation} where $T_0=vu, \theta=D_+v\times v^{-1}$. This system has partial solution $\alpha_0=-\beta_0=1$, which correspond to: $F_1=-u,F_2=v$. Transformation (\ref{31}) can be rewritten in terms of $T_0, \theta$ as: \begin{equation} D_+T_0=\theta T_0-T_0 \rig{\theta}{}\qquad D_-\theta=-T_0-\lef{T_0}{\:} \label{43} \end{equation} Taking now $D_-\alpha_0=\alpha_1\lef{T_0}{\:}+T_0\beta_1$ , for $\alpha_1, \beta_1$ we have: $$ D_+\alpha_1=\lef{\alpha}{}_0-\alpha_0+\theta\alpha_1+\alpha_1\lef{\theta}{} $$ $$ D_+\beta_1=\alpha_0-\rig{\alpha}{}_0+\rig{\theta}{}\beta_1+\beta_1\theta $$ For $\lef{\beta}{}_1=\alpha_1$ the second relation directly follows from the first one. For this case, acting on the equation for $\alpha_1$ with $D_-$ operator, we have: \begin{equation} \begin{array}{l} \beta_1=\rig{\alpha_1}{\!}\\ -D_+D_-\alpha_1=(\alpha_1+\lef{\alpha}{}_1)\lef{T_0}{2}-T_0(\alpha_1+\rig{\alpha_1}{\!})+ D_-\alpha_1\lef{\theta}{}-\theta D_-\alpha_1\\ \end{array} \label{44} \end{equation} This is the typical system for odd steps. Comparing it with (\ref{4}) we notice that the difference between those systems is that (\ref{44}) have not numerical partial solutions ($\alpha_1, \beta_1$ are Fermionic--like operators). However it is possible to continue reduction using decomposition $D_-\alpha_1=\alpha_2\lef{T_0}{2}+T_0\beta_2$. We have: $$ -D_+\alpha_2=\lef{\alpha}{}_1+\alpha_1-\theta\alpha_2+\alpha_2\lef{\theta}{2} $$ $$ -D_+\beta_2=-(\alpha_1+\rig{\alpha_1}{\!})-\rig{\theta}{}\beta_2+\beta_2\lef{\theta}{} $$ Taking $\lef{\beta}{}_2=-\alpha_2$, after usual simple calculations we will find: \begin{equation} \begin{array}{l} \beta_2=-\rig{\alpha_2}{\!}\\ D_+D_-\alpha_2=(\lef{\alpha}{}_2-\alpha_2)\lef{T_0}{3}+T_0(\alpha_2-\rig{\alpha_2}{\!})+ D_-\alpha_1\lef{\theta}{2}+\theta D_-\alpha_2\\ \end{array} \label{54} \end{equation} The partial solution of this system is: $\alpha_2=-\beta_2=1$; it correspond to the trivial system $F_1=au_x+bv_y, F_2=av_x+bv_y$. All systems received on even steps will be similar to (\ref{54}). Partial solution of each next system of that kind gives non--trivial, nonlinear evolution type system invariant with respect to the transformation (\ref{31}) (see $k=2$ example). By induction easily can be proved that for arbitrary $n=2k+1$ we will have: $$ D_-\alpha_{n-1}=\alpha_n\lef{T_0}{\:n}+T_0\rig{\alpha}{}_n $$ $$ D_-\alpha_n=\alpha_{n+1}\lef{T_0}{\:n+1}-T_0\rig{\alpha}{}_{n+1} $$ \begin{equation} \begin{array}{rr} D_+D_-\alpha_{n-1}=(\lef{\alpha}{}_{n-1}-\alpha_{n-1})\lef{T_0}{n}-T_0(\alpha_{n-1}- \rig{\alpha}{}_{n-1})+\\+D_-\alpha_{n-1}\lef{\theta}{n-1}+\theta D_-\alpha_{n-1} \end{array} \label{144} \end{equation} $$ -D_+D_-\alpha_n=(\lef{\alpha}{}_n+\alpha_n)\lef{T_0}{n+1}- T_0(\alpha_n+\rig{\alpha}{}_{n})+ D_-\alpha_n\lef{\theta}{n}\\-\theta D_-\alpha_n \label{145} $$ After this using $\alpha_{2k}=1$ partial solution of the system (\ref{144}) it is possible to construct the $k$--th term of hierarchy. One can prove using induction that the final result can be represented as: \begin{equation} \begin{array}{rcl} {\alpha_0}^{(k)}&=&(-1)^k \prod_{i=1}^{2k} \left(1-(-1)^iL_i exp\left[id_i+\sum_{i=k+1}^{2k} d_k\right]\right)\times \\&&\\ &&\times D_-^{-1}(T_0 D_-^{-1}(\rig{T_0}{\!}D_-^{-1} (...D_-{-1} \rig{T_0}{(n-1)}...)))\label{320} \end{array} \end{equation} The meaning of notations here is the same as in formula (\ref{20}). \subsection{Examples} \subsubsection*{k=0} $$ v_t=v\qquad u_t=-u $$ \subsubsection*{k=1} $$ v_t=v_x\qquad u_t=u_x $$ \subsubsection*{k=2} $$ v_t=v_{xx}-2D_-^{-1}(vu)_x D_+v-2D_-^{-1}(vD_+ u)\times v+ $$ $$ +2D_-^{-1} \left[vuD_-^{-1}(vu)_x +D_-^{-1}(vu)_x \times vu\right] $$ \vspace{1em} $$ u_t=-u_{xx}+D_+uD_-^{-1}(vu)_x -2uD_-^{-1}(D_+vu)_x- $$ $$ -2uD_-^{-1} \left[vuD_-^{-1}(vu)_x +D_-^{-1}(vu)_x \times vu\right] $$ \section{Conclusion} The main concrete result of the paper is the explicit form of quantum integrable systems (\ref{20}), (\ref{320}) in the mentioned above sense. It is interesting that the scheme of our calculations is similar to the computer program algorithm--there are many identical operations with possibility to interrupt them at arbitrary step. Obviously, in this scheme is coded the structure of the group of integrable mappings, more exactly one of the possible connections between the integrable system by itself and its symmetry equation. If it will be possible to translate it on the group--theoretical language, then we will be near to understand the integrable substitutions role and near to the classification theorem for them. It is well known that quantum integrable systems are closely connected with so--called "quantum" algebras \cite{16}. Moreover this object of mathematics in essential part was discovered and developed under the investigations of the integrable systems in quantum region. So it arise more wide, deep and interesting problem--to find the connection between the approach of this paper and sufficiently developed formalism of quantum algebras. The equations for Heisenberg operators, as it was mentioned in the introduction, are only one of the possible representations of the quantum picture. We hope that further investigations will find some bridge connecting quantum integrable mappings of the present paper with quantum algebras of the traditional approach. But now we are not ready and able to go so far and hope to return to this problem in the future publications. \section{Acknowledgments} The authors wish to thank the Russian Foundation for Fundamental Researches for partial support trough the Grant 95--01--00249
proofpile-arXiv_065-422
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} The presence of ubiquitous X-ray absorption in cooling flow (CF) clusters was first discovered by White et al. (1991) using the {\em Einstein} Solid State Spectrometer (SSS), and was later verified using {\em EXOSAT}, {\em ROSAT}, and {\em ASCA} (e.g. Allen et al. 1992; Fabian et al. 1994). The typical observed columns are a few $10^{20}$ to a few $10^{21}$~cm$^{-2}$. Searches were conducted at other wavelengths in order to verify the presence of the absorber and to understand its nature. Emission line constraints rule out absorbing gas at $T\sim 10^5-10^6$~K, and thus the absorbing gas must be mostly H~I and/or H$_2$. Extensive searches for H~I absorption using the 21~cm hyperfine structure line (e.g. Jaffe 1990; Dwarakanath, van Gorkom, \& Owen 1994) and searches for CO associated with H$_2$ (e.g. O'dea et al. 1994; Antonucci \& Barvainis 1994) yielded upper limits typically well below the X-ray columns. The 21~cm limits are linear with the electron excitation temperature, and may thus be subject to significant uncertainty. For example, in the case of NGC~1275, the central galaxy in the Perseus CF cluster, the X-ray column obtained by the {\em Einstein} SSS is $1.3^{+0.3}_{-0.3}\times 10^{21}$~cm$^{-2}$ (White et al. 1991), by {\em EXOSAT} $1.5^{+2.1}_{-0.9}\times 10^{21}$~cm$^{-2}$ (Allen et al. 1991), and by {\em ASCA} $3-4\times 10^{21}$~cm$^{-2}$ (Fabian et al. 1994). The 21~cm line, however, indicates an H~I column of only $2\times 10^{18}T_s$~cm$^{-2}$ (Jaffe 1990). In this paper we show that significantly tighter limits on $N_{\rm H~I}$ can be obtained using Ly$\alpha$. We first make a short comparison of the absorption properties of Ly$\alpha$ versus the 21~cm line. We then apply the results of curve of growth analysis for Ly$\alpha$ to show that the H~I column $\sim 10-20$~kpc away from the center of NGC~1275 is significantly smaller than indicated by the 21~cm line, and demonstrate that improved limits can be obtained with a higher quality UV spectrum. We end with a short discussion of the implications of the new upper limits on $N_{\rm H~I}$ on the nature of the X-ray absorber. \section{On Lyman $\alpha$ vs. 21 cm Absorption} In this section we compare the absorption properties of Lyman $\alpha$ versus the 21~cm line. In a two level atom the absorption cross section per atom is: \[ \sigma_{\nu}=\frac{\pi e^2}{m_ec}f_{12}\phi(\nu) f_{se}\frac{n_1}{n_1+n_2} \] where $f_{12}$ is the oscillator strength, and $\phi(\nu)$ is the line profile function (Voigt function for pure thermal broadening). The parameter $f_{se}$ is the correction for stimulated emission given by \[ f_{se}\equiv1-\frac{n_2}{n_1}\frac{g_1}{g_2} \] where $n_i, g_i$ are the population and degeneracy of level i. The level population is accurately described by the Boltzmann ratio since collisions dominate both excitations and deexcitaions, and thus \[ f_{se}=1-e^{-\frac{\Delta E}{kT}}, \] or $f_{se}\simeq \frac{\Delta E}{kT}$ for $\Delta E\ll kT$. The value of the line profile function at line center is \[ \phi(\nu_0)=\frac{1}{\sqrt{\pi}\Delta\nu_D}\] assuming a Gaussian line shape (i.e. thermal broadening), where $\Delta\nu_D=\nu_0\frac{b}{c}$ is the line width and $b=\sqrt{\frac{2kT}{m_p}}$ is the Doppler parameter (Rybicki \& Lightman 1979). Table 1 compares the values of the parameters discussed above for Ly$\alpha$ versus the 21~cm line. \begin{table} \caption{Ly$\alpha$ versus 21~cm absorption line parameters} \begin{center} \begin{tabular}{ccc} Parameter & Ly$\alpha$ & 21~cm\\ \tableline $f_{12}$ & 0.416 & $5.75\times 10^{-12}$ \\ $\phi(\nu_0)$ & $5.33\times 10^{-10}T^{-1/2}$ & $9.27\times 10^{-4}T^{-1/2}$\\ $f_{se}$ & $\simeq 1$ & $0.0682T^{-1}$\\ $n_2/n_1$ & $\ll 1$ & 3 \\ $ \sigma_{\nu_0}$ & $5.88\times10^{-12} T^{-1/2}$ & $2.41\times10^{-19} T^{-3/2}$ \\ \end{tabular} \end{center} \end{table} The ratio of line center absorption cross sections is therefore \[ \frac {\sigma_{\nu_0}({\rm Ly}\alpha)} {\sigma_{\nu_0}(21 {\rm cm})} =2.44\times 10^6 T, \] and since $T\ge 2.73$~K, Ly$\alpha$ is $10^7$ times more sensitive to H~I absorption than the 21~cm line, if both absorption lines are optically thin. Note that when $N_{\rm H~I}>2\times 10^{11}T^{1/2}$, the Ly$\alpha$ line becomes optically thick, and the absorption equivalent width $EW=\int (1-e^{-\tau_{\nu}})d\nu$ increases only as $\sqrt{\ln \tau}$. Figure 1 presents a comparison of the absorption profiles of the 21~cm line vs. Ly$\alpha$ for $10^{20}\ge N_{\rm H~I}\ge 10^{17}$~cm$^{-2}$. Note the large difference in absorption EW of the two lines. \begin{figure} \psfig{file=figcf1.ps,width=11.cm,angle=-90,silent=} \caption{A comparison of the absorption profiles of the 21~cm line (right) vs. Ly$\alpha$ (left) for $10^{17}\ge N_{\rm H~I}\ge 10^{20}$~cm$^{-2}$. Note the large difference in velocity scales in the two panels.} \end{figure} Jaffe (1990) measured in NGC~1275 21~cm absorption with $N_{\rm H~I}=2\times 10^{18}T$ and FWHM=477~km~s$^{-1}$ (i.e. $b=286$~km~s$^{-1}$). The expected Ly$\alpha$ absorption profile for $N_{\rm H~I}$ at various $T$ is displayed in Figure 2, \begin{figure} \psfig{file=figcf2.ps,width=11.cm,angle=-90,silent=} \caption{ The predicted vs. observed Ly$\alpha$ absorption profile. Left: The predicted absorption profile for different values of T with $N_{\rm H~I}=2\times 10^{18}T$ and $b=286$~km~s$^{-1}$, as measured by Jaffe (1990). Right: The Ly$\alpha$ spectrum observed by Johnstone \& Fabian (1995). Very little, if any, absorption is present in Ly$\alpha$.} \end{figure} indicating that for all $T$ one expects a very broad absorption trough with FWHM$\ge 2000$~km~s$^{-1}$. The Ly$\alpha$ region in NGC~1275 was observed by Johnstone \& Fabian (1995), and the observed spectrum is displayed in Figure 2 (velocity scale is relative to 5260~km~s$^{-1}$). Clearly, the absorption predicted based on the 21~cm $N_{\rm H~I}$ is not present. The small trough at the center of Ly$\alpha$ suggests absorption with EW$\sim 0.5$~\AA, or $120$~km~s$^{-1}$. Johnstone \& Fabian (1995) suggested that Ly$\alpha$ is double peaked, rather than absorbed, in which case the absorption EW would be $\ll 0.5$~\AA. Clearly, Ly$\alpha$ implies a much lower values for $N_{\rm H~I}$ than the 21~cm line. To obtain the $N_{\rm H~I}$ implied by Ly$\alpha$ one needs to calculate EW($N_{\rm H~I}$), i.e. use the standard ``curve of growth'' analysis. Figure 3 displays on the left hand side the Ly$\alpha$ EW versus $N_{\rm H~I}$ for various values of the $b$ parameter (assuming a Gaussian velocity distribution). The EW increases linearly with $N_{\rm H~I}$ when the line is optically thin, saturating to EW$\propto \sqrt{\ln N_{\rm H~I}}$ when the line becomes optically thick, and recovering back to EW$\propto \sqrt{N_{\rm H~I}}$ when the Lorenztian wings dominate the absorption (`damped' absorption). The observed absorption EW of 120~km~s$^{-1}$ translates to $10^{14}\ge N_{\rm H~I}\ge 4\times 10^{17}$~cm$^{-2}$, where the upper limit is obtain if $b<10$~km~s$^{-1}$, and the lower limit is obtained if $b>50$~km~s$^{-1}$. The right hand side curves in Figure 3 represent the curves of growth for 21~cm absorption by H~I at $T=10$~K. These curves are identical to those for Ly$\alpha$ absorption, but shifted by a factor of $2.4\times 10^7$ to the right hand side. The observed 21~cm absorption EW translates to $N_{\rm H~I}=2\times 10^{19}$~cm$^{-2}$ for a reasonable lower limit of $T=10$~K. The largest column allowed by Ly$\alpha$ is therefore $\sim 50$ times smaller than indicated by the 21~cm line. \begin{figure} \psfig{file=figcf3.ps,width=13.0cm,silent=} \caption{The Ly$\alpha$ and the 21~cm curves of growth for different velocity dispersions ($b$ parameters). The horizontal dashed lines indicate the observed absorption EWs, and the vertical dashed lines indicate the derived limits on $N_{\rm H~I}$. The largest column allowed by Ly$\alpha$ is $\sim 50$ times smaller than indicated by the 21~cm line.} \end{figure} \section{How can the 21~cm and Ly$\alpha$ columns be reconciled?} The apparent contradiction between the 21~cm and the Ly$\alpha$ columns can be understood if the H~I column is highly non-uniform, and the spatial distributions of the background 21~cm and Ly$\alpha$ emission are different. There is observational evidence that both these effects are present in NGC~1275. According to the 21~cm continuum map presented by Jaffe (1990) most of the continuum originates within $\pm 20''$ of the center. Fabian, Nulsen, \& Arnaud (1984) discovered with IUE that Ly$\alpha$ is also extended on $\sim 10''$ scale, and the recent HUT observations by Van Dyke Dixon, Davidsen, \& Ferguson (1996) indicate that Ly$\alpha$ emission of comparable surface brightness extends out to $\sim 60''$ from the center. Evidence for spatially non-uniform absorption is seen on much smaller scales in the VLBA observations of Walker et al. (1994) and Vermeulen et al. (1994), who discovered a free-free absorbed counter jet to the north of the nucleus. The counter jet is most likely seen through a large column disk of relatively cold gas close to the center of NGC~1275 (Levinson et al. 1995), while the line of sight to the southren jet is clear. The large $N_{\rm H~I}$ indicated by the 21~cm line most likely resides on scales smaller than the $\sim 10-20$~kpc scale of the Ly$\alpha$ emitting filaments. The low $N_{\rm H~I}$ indicated by Ly$\alpha$ provides a constraint for the $\sim 100-200$~kpc scale absorber indicated by the X-ray observations. \section{How can the X-ray and Ly$\alpha$ columns be reconciled?} The various X-ray telescopes mentioned above indicate an excess absorbing column of $(1.5-4)\times 10^{21}$~cm$^{-2}$, and a covering factor close to unity. Such a column becomes optically thick at $E<0.6-1$~keV. At this energy range O is the dominant absorber (e.g. Morrison \& McCammon 1983). Thus, the X-ray spectra merely indicate the presence of an O column of $(1.3-3.4)\times 10^{18}$~cm$^{-2}$. The O X-ray absorption is done by the inner K shell electrons, thus the ionization state of the O can be anywhere from O~I to O~VII. The X-ray and Ly$\alpha$ constraints imply that whatever is producing the absorption on the $\sim 100$~kpc scale in the Perseus cooling flow cluster must have $3<N_{\rm O}/N_{\rm H~I}<3\times 10^4$, i.e. it is drastically different from a neutral, solar abundance absorber, where $N_{\rm O}/N_{\rm H~I}=8.5\times 10^{-4}$. If the absorber has roughly solar abundance then H must be highly ionized with $2.5\times 10^{-8}<N_{\rm H~I}/N_{\rm H}<2.5\times 10^{-4}$. Can H be so highly ionized? The available ionizing flux is far too low for significant photoionization. Collisional ionization requires $5\times 10^6>T>5\times 10^4$~K, where the upper limit prevents O from being too highly ionized. However, this temperature range is excluded based on the absence of significant line emission (e.g. Voit \& Donahue 1995). It thus appears that the required ionization state of H is ruled out. The above constraints on $T$ assume equilibrium ionization states. It remains to be studied whether plausible deviations from ionization equilibrium can significantly affect the ionization state of H. Another possibility is that the absorber has practically no H. This would be the case if the absorption originates in O which resides in dust grains embeded in hot gas. However, the dust sputtering time scales appear too short to explain the absorption in the inner parts of clusters (Dwek et al. 1990; Voit \& Donahue 1995). Could most of the H be in molecular form? CO emission was detected in the Perseus CF (e.g. Braine et al. 1995) indicating that some of the H is indeed in molecular form. However, the H~I/H$_2$ fraction needs to be $<2.5\times 10^{-4}$, while theoretical calculations (Ferland et al. 1994) indicate that most of H ($>80$\%) would be in atomic form, even in extremely cold clouds embeded in CFs. It therefore appears that there is no satisfactory model which explains the X-ray absorption together with the new tight limits on $N_{\rm H~I}$ obtained from Ly$\alpha$. \section{Future perspectives} Given the difficulty in finding a plausible explanation for the X-ray absorption, it is crucial to verify that this absorption is indeed real. This can be achieved by the detection of an O bound-free K edge at the CF cluster redshift (see Sarazin, these proceedings). The edge energy will also indicate the ionization state of the absorber. This can be achieved with next generation high resolution X-ray telescopes. On a shorter time scale, significantly better constraints on $N_{\rm H~I}$ in NGC~1275 can be obtained with HST. Currently, the actual value of $N_{\rm H~I}$ in NGC~1275 is uncertain by nearly 4 orders of magnitude, and a much more accurate determination can be achieved through a higher resolution UV spectrum, as demonstrated in Figure 4. \begin{figure} \psfig{file=figcf4.ps,width=11.cm,angle=-90,silent=} \caption{The expected Ly$\alpha$ profile in NGC~1275 at a high spectral resolution ($\sim 10$~km~s$^{-1}$). Left: Emission + absorption profiles. Right: Blowout of the absorption profile. Such a spectrum would allow a rather accurate determination of $N_{\rm H~I}$. If the H~I gas has a very low velocity dispersion the absorption will go black, and if the velocity dispersion is high the absorption profile will remain shallow.} \end{figure} The method described here can be extended to many more CF clusters. All CF clusters have a large cD galaxy at their center, and these galaxies tend to have a power-law continuum source at their center with significant UV emission which can be used for the detection of Ly$\alpha$ absorption. In addition, significant Ly$\alpha$ emission most likely originates from the emission line filaments present in all CF clusters. For example, Hu (1992) observed 10 CF cluster with the IUE, detecting significant Ly$\alpha$ emission in 7 clusters, thus demonstrating that the method described here can be applied to most CF clusters. The disadvantage of using the central cD galaxy is that if absorption is detected it may be produced by gas local to the cD or the emission line filaments, rather than the large scale absorber. The lack of significant absorption can, however, be used to place stringent limits on the nature of the X-ray absorber. \acknowledgments This research was partly supported by the E. and J. Bishop research fund, and by the Milton and Lillian Edwards academic lectureship fund. The UV spectrum of NGC~1275 was generously provided by R. M. Johnstone.
proofpile-arXiv_065-423
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} The precision data collected to date have confirmed the Standard Model to be a good description of physics below the electroweak scale \cite{Schaile}. Despite of its great success, there are many reasons to believe that some kind of new physics must exist. On the other hand, the non-abelian structure of the gauge boson self-couplings is still poorly tested and one of the most sensitive probes for new physics is provided by the trilinear gauge boson couplings (TGC) \cite{TGC}. Many studies have been devoted to the $WW\gamma$ and $WWZ$ couplings. At hadron colliders and $e^+e^-$ colliders, the present bounds (Tevatron \cite{Errede}) and prospects (LHC, LEP2 and NLC \cite{TGC,LEP2}) are mostly based on diboson production ($WW$, $W\gamma$ and $WZ$). In $ep$ collisions, HERA could provide further information analyzing single $W$ production ($ep\to eWX$ \cite{ABZ}) and radiative charged current scattering ($ep\to\nu\gamma X$ \cite{hubert}). There is also some literature on $WW\gamma$ couplings in $W$-pair production at future very high energy photon colliders (bremsstrahlung photons in peripheral heavy ion collisions \cite{HIC} and Compton backscattered laser beams \cite{gg}). Only recently, attention has been paid to the $Z\gamma Z$, $Z\gamma\g$ and $ZZZ$ couplings. There is a detailed analysis of $Z\gamma V$ couplings ($V=\gamma,Z$) for hadron colliders in \cite{BB}. CDF \cite{CDF} and D\O\ \cite{D0} have obtained bounds on the $Z\gamma Z$ and $Z\gamma\g$ anomalous couplings, while L3 has studied only the first ones \cite{L3}. Studies on the sensitivities to these vertices in future $e^+e^-$ colliders, LEP2 \cite{LEP2} and NLC \cite{Boudjema}, have been performed during the last years. Some proposals have been made to probe these neutral boson gauge couplings at future photon colliders in $e\gamma\to Ze$ \cite{eg}. In this work we study the prospects for measuring the TGC in the process $ep\to e\gamma X$. In particular, we will concentrate on the $Z\gamma\g$ couplings, which can be more stringently bounded than the $Z\gamma Z$ ones for this process. In Section 2, we present the TGC. The next section deals with the different contributions to the process $ep\to e\gamma X$ and the cuts and methods we have employed in our analysis. Section 4 contains our results for the Standard Model total cross section and distributions and the estimates of the sensitivity of these quantities to the presence of anomalous couplings. Finally, in the last section we present our conclusions. \section{Phenomenological parametrization of the neutral TGC} A convenient way to study deviations from the standard model predictions consists of considering the most general lagrangian compatible with Lorentz invariance, the electromagnetic U(1) gauge symmetry, and other possible gauge symmetries. For the trilinear $Z\gamma V$ couplings ($V=\gamma,Z)$ the most general vertex function invariant under Lorentz and electromagnetic gauge transformations can be described in terms of four independent dimensionless form factors \cite{hagiwara}, denoted by $h^V_i$, i=1,2,3,4: \begin{eqnarray} \Gamma^{\a\b\mu}_{Z\gamma V} (q_1,q_2,p)=\frac{f(V)}{M^2_Z} \{ h^V_1 (q^\mu_2 g^{\a\b} - q^\a_2 g^{\mu\b}) +\frac{h^V_2}{M^2_Z} p^\a (p\cdot q_2g^{\mu\b}-q^\mu_2 p^\b) \nonumber \\ +h^V_3 \varepsilon^{\mu\a\b\r}q_{2_\r} +\frac{h^V_4}{M^2_Z}p^\a\varepsilon^{\mu\b\r\sigma}p_\r q_{2_\sigma} \}. \hspace{3cm} \label{vertex} \end{eqnarray} Terms proportional to $p^\mu$, $q^\a_1$ and $q^\b_2$ are omitted as long as the scalar components of all three vector bosons can be neglected (whenever they couple to almost massless fermions) or they are zero (on-shell condition for $Z$ or U(1) gauge boson character of the photon). The overall factor, $f(V)$, is $p^2-q^2_1$ for $Z\gamma Z$ or $p^2$ for $Z\gamma\g$ and is a result of Bose symmetry and electromagnetic gauge invariance. These latter constraints reduce the familiar seven form factors of the most general $WWV$ vertex to only these four for the $Z\gamma V$ vertex. There still remains a global factor that can be fixed, without loss of generality, to $g_{Z\gamma Z}=g_{Z\gamma\g}=e$. Combinations of $h^V_3 (h^V_1)$ and $h^V_4 (h^V_2)$ correspond to electric (magnetic) dipole and magnetic (electric) quadrupole transition moments in the static limit. All the terms are $C$-odd. The terms proportional to $h^V_1$ and $h^V_2$ are $CP$-odd while the other two are $CP$-even. All the form factors are zero at tree level in the Standard Model. At the one-loop level, only the $CP$-conserving $h^V_3$ and $h^V_4$ are nonzero \cite{barroso} but too small (${\cal O}(\a/\pi$)) to lead to any observable effect at any present or planned experiment. However, larger effects might appear in theories or models beyond the Standard Model, for instance when the gauge bosons are composite objects \cite{composite}. This is a purely phenomenological, model independent parametrization. Tree-level unitarity restricts the $Z\gamma V$ to the Standard Model values at asympotically high energies \cite{unitarity}. This implies that the couplings $h^V_i$ have to be described by form factors $h^V_i(q^2_1,q^2_2,p^2)$ which vanish when $q^2_1$, $q^2_2$ or $p^2$ become large. In hadron colliders, large values of $p^2=\hat{s}$ come into play and the energy dependence has to be taken into account, including unknown dumping factors \cite{BB}. A scale dependence appears as an additional parameter (the scale of new physics, $\L$). Alternatively, one could introduce a set of operators invariant under SU(2)$\times$U(1) involving the gauge bosons and/or additional would-be-Goldstone bosons and the physical Higgs. Depending on the new physics dynamics, operators with dimension $d$ could be generated at the scale $\L$, with a strength which is generally suppressed by factors like $(M_W/\L)^{d-4}$ or $(\sqrt{s}/\L)^{d-4}$ \cite{NPscale}. It can be shown that $h^V_1$ and $h^V_3$ receive contributions from operators of dimension $\ge 6$ and $h^V_2$ and $h^V_4$ from operators of dimension $\ge 8$. Unlike hadron colliders, in $ep\to e\gamma X$ at HERA energies, we can ignore the dependence of the form factors on the scale. On the other hand, the anomalous couplings are tested in a different kinematical region, which makes their study in this process complementary to the ones performed at hadron and lepton colliders. \section{The process $ep\to e\gamma X$} The process under study is $ep\to e\gamma X$, which is described in the parton model by the radiative neutral current electron-quark and electron-antiquark scattering, \begin{equation} \label{process} e^- \ \stackrel{(-)}{q} \to e^- \ \stackrel{(-)}{q} \ \gamma . \end{equation} There are eight Feynman diagrams contributing to this process in the Standard Model and three additional ones if one includes anomalous vertices: one extra diagram for the $Z\gamma Z$ vertex and two for the $Z\gamma\g$ vertex (Fig. \ref{feyndiag}). \bfi{htb} \begin{center} \bigphotons \bpi{35000}{21000} \put(4000,8000){(a)} \put(200,17000){\vector(1,0){1300}} \put(1500,17000){\vector(1,0){3900}} \put(5400,17000){\line(1,0){2600}} \drawline\photon[\S\REG](2800,17000)[5] \put(200,\pbacky){\vector(1,0){1300}} \put(1500,\pbacky){\vector(1,0){2600}} \put(4100,\pbacky){\vector(1,0){2600}} \put(6700,\pbacky){\line(1,0){1300}} \put(0,13000){$q$} \put(8200,13000){$q$} \put(3300,\pmidy){$\gamma,Z$} \drawline\photon[\SE\FLIPPED](4900,\pbacky)[4] \put(0,18000){$e$} \put(8200,18000){$e$} \put(8200,\pbacky){$\gamma$} \put(13000,8000){(b)} \put(9500,17000){\vector(1,0){1300}} \put(10800,17000){\vector(1,0){2600}} \put(13400,17000){\vector(1,0){2600}} \put(16000,17000){\line(1,0){1300}} \drawline\photon[\S\REG](12100,17000)[5] \put(9500,\pbacky){\vector(1,0){1300}} \put(10800,\pbacky){\vector(1,0){3900}} \put(14700,\pbacky){\line(1,0){2600}} \drawline\photon[\NE\FLIPPED](14200,17000)[4] \put(22000,8000){(c)} \put(18500,17000){\vector(1,0){3250}} \put(21750,17000){\vector(1,0){3250}} \put(25000,17000){\line(1,0){1300}} \drawline\photon[\S\REG](23700,17000)[5] \put(18500,\pbacky){\vector(1,0){1300}} \put(19800,\pbacky){\vector(1,0){2600}} \put(22400,\pbacky){\vector(1,0){2600}} \put(25000,\pbacky){\line(1,0){1300}} \drawline\photon[\SE\FLIPPED](21100,\pbacky)[4] \put(31000,8000){(d)} \put(27500,17000){\vector(1,0){1300}} \put(28800,17000){\vector(1,0){2600}} \put(31400,17000){\vector(1,0){2600}} \put(34000,17000){\line(1,0){1300}} \drawline\photon[\S\REG](32700,17000)[5] \put(27500,\pbacky){\vector(1,0){3250}} \put(30750,\pbacky){\vector(1,0){3250}} \put(33900,\pbacky){\line(1,0){1300}} \drawline\photon[\NE\FLIPPED](30100,17000)[4] \put(17800,0){(e)} \put(17100,5500){$\gamma,Z$} \put(17100,3000){$\gamma,Z$} \put(14000,7000){\vector(1,0){1300}} \put(15300,7000){\vector(1,0){3900}} \put(19200,7000){\line(1,0){2600}} \drawline\photon[\S\REG](16600,7000)[5] \put(16750,\pmidy){\circle*{500}} \put(14000,\pbacky){\vector(1,0){1300}} \put(15300,\pbacky){\vector(1,0){3900}} \put(19200,\pbacky){\line(1,0){2600}} \drawline\photon[\E\REG](16750,\pmidy)[5] \put(22300,\pbacky){$\gamma$} \end{picture} \end{center} \caption{\it Feynman diagrams for the process $e^- q \to e^- q \gamma$. \label{feyndiag}} \end{figure} Diagrams with $\gamma$ exchanged in the t-channel are dominant. Nevertheless, we consider the whole set of diagrams in the calculation. On the other side, u-channel fermion exchange poles appear, in the limit of massless quarks and electrons (diagrams (c) and (d)). Since the anomalous diagrams (e) do not present such infrared or collinear singularities, it seems appropriate to avoid almost on-shell photons exchanged and fermion poles by cutting the transverse momenta of the final fermions (electron and jet) to enhance the signal from anomalous vertices. Due to the suppression factor coming from $Z$ propagator, the anomalous diagrams are more sensitive to $Z\gamma\g$ than to $Z\gamma Z$ vertices. In the following we will focus our attention on the former. The basic variables of the parton level process are five. A suitable choice is: $E_\gamma$ (energy of the final photon), $\cos\th_\gamma$, $\cos\th_{q'}$ (cosines of the polar angles of the photon and the scattered quark defined with respect to the proton direction), $\phi$ (the angle between the transverse momenta of the photon and the scattered quark in a plane perpendicular to the beam), and a trivial azimuthal angle that is integrated out (unpolarized beams). All the variables are referred to the laboratory frame. One needs an extra variable, the Bjorken-x, to connect the partonic process with the $ep$ process. The phase space integration over these six variables is carried out by {\tt VEGAS} \cite{VEGAS} and has been cross-checked with the {\tt RAMBO} subroutine \cite{RAMBO}. We adopt two kinds of event cuts to constrain conveniently the phase space: \begin{itemize} \item {\em Acceptance and isolation} cuts. The former are to exclude phase space regions which are not accessible to the detector, because of angular or efficiency limitations:\footnote{The threshold for the transverse momentum of the scattered quark ensures that its kinematics can be described in terms of a jet.} \begin{eqnarray} \label{cut1} 8^o < \theta_e,\ \theta_\gamma,\ \theta_{\rm jet} < 172^o; \nonumber\\ E_e, \ E_\gamma, \ p^{\rm q'}_{\rm T} > 10 \ {\rm GeV}. \end{eqnarray} The latter keep the final photon well separated from both the final electron and the jet: \begin{eqnarray} \label{cut2} \cos \langle \gamma,e \rangle < 0.9; \nonumber\\ R > 1.5, \end{eqnarray} where $R\equiv\sqrt{\Delta\eta^2+\phi^2}$ is the separation between the photon and the jet in the rapidity-azimuthal plane, and $\langle \gamma,e \rangle$ is the angle between the photon and the scattered electron. \item Cuts for {\em intrinsic background suppression}. They consist of strengthening some of the previous cuts or adding new ones to enhance the signal of the anomalous diagrams against the Standard Model background. \end{itemize} We have developed a Monte Carlo program for the simulation of the process $ep\to e\gamma X$ where $X$ is the remnant of the proton plus one jet formed by the scattered quark of the subprocess (\ref{process}). It includes the Standard Model helicitity amplitudes computed using the {\tt HELAS} subroutines \cite{HELAS}. We added new code to account for the anomalous diagrams. The squares of these anomalous amplitudes have been cross-checked with their analytical expressions computed using {\tt FORM} \cite{FORM}. For the parton distribution functions, we employ both the set 1 of Duke-Owens' parametrizations \cite{DO} and the modified MRS(A) parametrizations \cite{MRS}, with the scale chosen to be the hadronic momentum transfer. As inputs, we use the beam energies $E_e=30$ GeV and $E_p=820$ GeV, the $Z$ mass $M_Z=91.187$ GeV, the weak angle $\sin^2_W=0.2315$ \cite{PDB} and the fine structure constant $\a=1/128$. A more correct choice would be the running fine structure constant with $Q^2$ as the argument. However, as we are interested in large $Q^2$ events, the value $\a(M^2_Z)$ is accurate enough for our purposes. We consider only the first and second generations of quarks, assumed to be massless. We start by applying the cuts (\ref{cut1}) and (\ref{cut2}) and examining the contribution to a set of observables of the Standard Model and the anomalous diagrams, separately. Next, we select one observable such that, when a cut on it is performed, only Standard Model events are mostly eliminated. The procedure is repeated with this new cut built in. After several runs, adding new cuts, the ratio standard/anomalous cross sections is reduced and hence the sensitivity to anomalous couplings is improved. \section{Results} \subsection{Observables} The total cross section of $ep\to e\gamma X$ can be written as \begin{equation} \sigma=\sigma_{{\rm SM}} + \sum_{i} \t_i \cdot h^\gamma_i + \sum_{i}\sigma_i\cdot (h^\gamma_i)^2 + \sigma_{12} \cdot h^\gamma_1 h^\gamma_2 + \sigma_{34} \cdot h^\gamma_3 h^\gamma_4. \end{equation} \bfi{htb} \setlength{\unitlength}{1cm} \bpi{8}{7} \epsfxsize=11cm \put(-1,-4){\epsfbox{eng_acciso.ps}} \end{picture} \bpi{8}{7} \epsfxsize=11cm \put(0.,-4){\epsfbox{ptg_acciso.ps}} \end{picture} \bpi{8}{6} \epsfxsize=11cm \put(-1,-5){\epsfbox{angge_acciso.ps}} \end{picture} \bpi{8}{6} \epsfxsize=11cm \put(0.,-5){\epsfbox{anggj_acciso.ps}} \end{picture} \bpi{8}{6} \epsfxsize=11cm \put(-1,-5){\epsfbox{angej_acciso.ps}} \end{picture} \bpi{8}{7} \epsfxsize=11cm \put(0.,-5){\epsfbox{q2e_acciso.ps}} \end{picture} \caption{\it Differential cross sections (pb) for the process $ep\to e\gamma X$ at HERA, with only acceptance and isolation cuts. The solid line is the Standard Model contribution and the dash (dot-dash) line correspond to 10000 times the $\sigma_1$ ($\sigma_2$) anomalous contributions.\label{A}} \end{figure} The forthcoming results are obtained using the MRS'95 pa\-ra\-me\-tri\-za\-tion of the parton densities\footnote{The values change $\sim 10$\% when using the (old) Duke-Owens' structure functions.} \cite{MRS}. The linear terms of the $P$-violating couplings $h^\gamma_3$ and $h^\gamma_4$ are negligible, as they mostly arise from the interference of standard model diagrams with photon exchange ($P$-even) and anomalous $P$-odd diagrams ($\t_3\simeq \t_4\simeq 0$). Moreover, anomalous diagrams with different $P$ do not interfere either. On the other hand, the quadratic terms proportional to $(h^\gamma_1)^2$ and $(h^\gamma_3)^2$ have identical expressions, and the same for $h^\gamma_2$ and $h^\gamma_4$ ($\sigma_1=\sigma_3$, $\sigma_2=\sigma_4$). Only the linear terms make their bounds different. The interference terms $\sigma_{12}$ and $\sigma_{34}$ are also identical. \bfi{htb} \setlength{\unitlength}{1cm} \bpi{8}{7} \epsfxsize=11cm \put(-1,-4){\epsfbox{eng_bkgsup.ps}} \end{picture} \bpi{8}{7} \epsfxsize=11cm \put(0.,-4){\epsfbox{ptg_bkgsup.ps}} \end{picture} \bpi{8}{6} \epsfxsize=11cm \put(-1,-5){\epsfbox{angge_bkgsup.ps}} \end{picture} \bpi{8}{6} \epsfxsize=11cm \put(0.,-5){\epsfbox{anggj_bkgsup.ps}} \end{picture} \bpi{8}{6} \epsfxsize=11cm \put(-1,-5){\epsfbox{angej_bkgsup.ps}} \end{picture} \bpi{8}{7} \epsfxsize=11cm \put(0.,-5){\epsfbox{q2e_bkgsup.ps}} \end{picture} \caption{\it Differential cross sections (pb) for the process $ep\to e\gamma X$ at HERA, after intrinsic background suppression. The solid line is the Standard Model contribution and the dash (dot-dash) line correspond to 500 times the $\sigma_1$ ($\sigma_2$) anomalous contributions.\label{B}} \end{figure} We have analyzed the distributions of more than twenty observables in the laboratory frame, including the energies, transverse momenta and angular distributions of the jet, the photon and the final electron, as well as their spatial, polar and azimuthal separations. Also the bjorken-x, the leptonic and hadronic momenta transfer and other fractional energies are considered. The process of intrinsic background suppression is illustrated by comparing Figures \ref{A} and \ref{B}. For simplicity, only the most interesting variables are shown: the energy $E(\gamma)$ and transverse momentum $p_T(\gamma)$ of the photon; the angles between the photon and the scattered electron $\langle \gamma,e \rangle$, the photon and the jet $\langle \gamma,j \rangle$, and the scattered electron and the jet $\langle e,j \rangle$; and the leptonic momentum transfer $Q^2(e)$. In Fig.~\ref{A}, these variables are plotted with only acceptance and isolation cuts implemented. All of them share the property of disposing of a range where any anomalous effect is negligible, whereas the contribution to the total SM cross section is large. The set of cuts listed below were added to reach eventually the distributions of Fig.~\ref{B}: \begin{itemize} \item The main contribution to the Standard Model cross section comes from soft photons with very low transverse momentum. The following cuts suppress a 97$\%$ of these events, without hardly affecting the anomalous diagrams which, conversely, enfavour high energy photons: \begin{eqnarray} E_\gamma > 30 \ {\rm GeV} \nonumber \\ p^\gamma_T > 20 \ {\rm GeV} \label{cut3} \end{eqnarray} \item Another remarkable feature of anomalous diagrams is the very different typical momentum transfers. Let's concentrate on the leptonic momentum transfer, $Q^2_e=-(p'_e-p_e)^2$. The phase space enhances high $Q^2_e$, while the photon propagator of the Standard Model diagrams prefer low values (above the threshold for electron detectability, $Q^2_e>5.8$~GeV$^2$, with our required minimum energy and angle). On the contrary, the anomalous diagrams have always a $Z$ propagator which introduces a suppression factor of the order of $Q^2_e/M^2_Z$ and makes irrelevant the $Q^2_e$ dependence, which is only determined by the phase space. As a consequence, the following cut looks appropriate, \begin{equation} Q^2_e > 1000 \ {\rm GeV}^2 \label{cut4} \end{equation} \end{itemize} It is important to notice at this point why usual form factors for the anomalous couplings can be neglected at HERA. For our process, these form factors should be proportional to $1/(1+Q^2/\L^2)^n$. With the scale of new physics $\L=500$~GeV to 1~TeV, these factors can be taken to be one. This is not the case in lepton or hadron high energy colliders where the diboson production in the s-channel needs dumping factors $1/(1+\hat{s}/\L^2)^n$. The total cross section for the Standard Model with acceptance and isolation cuts is $\sigma_{\rm SM}=21.38$~pb and is reduced to 0.37~pb when all the cuts are applied, while the quadratic contributions only change from $\sigma_1=2\times10^{-3}$~pb, $\sigma_2=1.12\times10^{-3}$~pb to $\sigma_1=1.58\times10^{-3}$~pb, $\sigma_2=1.05\times10^{-3}$~pb. The linear terms are of importance and change from $\t_1=1.18\times10^{-2}$~pb, $\t_2=1.27\times10^{-3}$~pb to $\t_1=7.13\times10^{-3}$~pb, $\t_2=1.26\times10^{-3}$~pb. Finally, the interference term $\sigma_{12}=1.87\times10^{-3}$~pb changes to $\sigma_{12}=1.71\times10^{-3}$~pb. The typical Standard Model events consist of soft and low-$p_T$ photons mostly backwards, tending to go in the same direction of the scattered electrons (part of them are emitted by the hadronic current in the forward direction), close to the required angular separation ($\sim 30^o$). The low-$p_T$ jet goes opposite to both the photon and the scattered electron, also in the transverse plane. On the contrary, the anomalous events have not so soft and high-$p_T$ photons, concentrated in the forward region as it the case for the scattered electron and the jet. \subsection{Sensitivity to anomalous couplings} In order to estimate the sensitivity to anomalous couplings, we consider the $\chi^2$ function. One can define the $\chi^2$, which is related to the likelihood function ${\cal L}$, as \begin{equation} \label{chi2} \chi^2\equiv-2\ln{\cal L}= 2 L \displaystyle\left(\sigma^{th}-\sigma^{o}+\sigma^{o} \ln\displaystyle\frac{\sigma^{o}}{\sigma^{th}}\right) \simeq L \displaystyle\dis\frac{(\sigma^{th}-\sigma^{o})^2}{\sigma^{o}}, \end{equation} where $L=N^{th}/\sigma^{th}=N^o/\sigma^o$ is the integrated luminosity and $N^{th}$ ($N^o$) is the number of theoretical (observed) events. The last line of (\ref{chi2}) is a useful and familiar approximation, only valid when $\mid \sigma^{th}-\sigma^o \mid/ \sigma^o \ll 1$. This function is a measure of the probability that statistical fluctuations can make undistinguisable the observed and the predicted number of events, that is the Standard Model prediction. The well known $\chi^2$-CL curve allows us to determine the corresponding confidence level (CL). We establish bounds on the anomalous couplings by fixing a certain $\chi^2=\d^2$ and allowing for the $h^\gamma_i$ values to vary, $N^o=N^o(h^\gamma_i)$. The parameter $\d$ is often referred as the number of standard deviations or `sigmas'. A $95\%$ CL corresponds to almost two sigmas ($\d=1.96$). When $\sigma \simeq \sigma_{{\rm SM}} + (h^\gamma_i)^2 \sigma_i$ (case of the $CP$-odd terms) and the anomalous contribution is small enough, the upper limits present some useful, approximate scaling properties, with the luminosity, \begin{equation} h^\gamma_i (L')\simeq\sqrt[4]{\frac{L}{L'}} \ h^\gamma_i (L). \end{equation} A brief comment on the interpretation of the results is in order. As the cross section grows with $h^\gamma_i$, in the relevant range of values, the $N^o$ upper limits can be regarded as the lowest number of measured events that would discard the Standard Model, or the largest values of $h^\gamma_i$ that could be bounded if no effect is observed, with the given CL. This procedure approaches the method of upper limits for Poisson processes when the number of events is large ($\mathrel{\rlap{\lower4pt\hbox{\hskip1pt$\sim$} 10$). \bfi{htb} \setlength{\unitlength}{1cm} \bpi{8}{8} \epsfxsize=12cm \put(3.35,4.245){+} \put(-2.5,-1.5){\epsfbox{conh1h2.nogrid.ps}} \end{picture} \bpi{8}{8} \epsfxsize=12cm \put(4.1,4.245){+} \put(-1.75,-1.5){\epsfbox{conh3h4.nogrid.ps}} \end{picture} \caption{\it Limit contours for $Z\gamma\g$ couplings at HERA with an integrated luminosity of 10, 100, 250, 1000 pb$^{-1}$ and a 95\% CL.\label{contour}} \end{figure} In Fig. \ref{contour} the sensitivities for different luminosities are shown. Unfortunately, HERA cannot compete with Tevatron, whose best bounds, reported by the D\O\ collaboration \cite{D0}, are \begin{eqnarray} |h^\gamma_1|, \ |h^\gamma_3| &<& 1.9 \ (3.1), \nonumber \\ |h^\gamma_2|, \ |h^\gamma_4| &<& 0.5 \ (0.8). \end{eqnarray} For the first value it was assumed that only one anomalous coupling contributes (`axial limits') and for the second there are two couplings contributing (`correlated limits'). Our results are summarized in Table \ref{table}. \begin{table} \begin{center} \begin{tabular}{|c|r|r|r|r|r|r|r|r|} \hline HERA & \multicolumn{2}{c|}{10 pb$^{-1}$} & \multicolumn{2}{c|}{100 pb$^{-1}$} & \multicolumn{2}{c|}{250 pb$^{-1}$} & \multicolumn{2}{c|}{1 fb$^{-1}$} \\ \hline \hline $h^\gamma_1$ & -19.0 & 14.5 & -11.5 & 7.0 & -9.5 & 5.5 & -8.0 & 3.5 \\ & -26.0 & 19.5 & -16.0 & 9.5 & -14.0 & 7.0 & -11.5 & 4.5 \\ \hline $h^\gamma_2$ & -21.5 & 20.0 & -12.0 & 10.0 & - 9.5 & 8.0 & -7.0 & 6.0 \\ & -26.0 & 30.0 & -13.0 & 18.0 & -10.0 & 15.0 & - 7.5 & 12.0 \\ \hline $h^\gamma_3$ & -17.0 & 17.0 & -9.0 & 9.0 & -7.5 & 7.5 & -5.5 & 5.5 \\ & -22.5 & 22.5 & -12.0 & 12.0 & -10.0 & 10.0 & -7.0 & 7.0 \\ \hline $h^\gamma_4$ & -20.5 & 20.5 & -11.0 & 11.0 & -8.5 & 8.5 & -6.0 & 6.0 \\ & -27.5 & 27.5 & -14.5 & 14.5 & -12.0 & 12.0 & -8.5 & 8.5 \\ \hline \end{tabular} \end{center} \caption{\it Axial and correlated limits for the $Z\gamma\g$ anomalous couplings at HERA with different integrated luminosities and $95\%$ CL. \label{table}} \end{table} The origin of so poor results is the fact that, unlike diboson production at hadron or $e^+e^-$ colliders, the anomalous diagrams of $ep\to e\gamma X$ have a $Z$ propagator decreasing their effect. The process $ep\to eZX$ avoids this problem thanks to the absence of these propagators: the Standard Model cross section is similar to the anomalous one but, as a drawback, they are of the order of femtobarns. \section{Summary and conclusions} The radiative neutral current process $ep\to e\gamma X$ at HERA has been studied. Realistic cuts have been applied in order to observe a clean signal consisting of detectable and well separated electron, photon and jet. The possibility of testing the trilinear neutral gauge boson couplings in this process has also been explored. The $Z\gamma Z$ couplings are very suppressed by two $Z$ propagators. Only the $Z\gamma \gamma$ couplings have been considered. A Monte Carlo program has been developed to account for such anomalous vertex and further cuts have been implemented to improve the sensitivity to this source of new physics. Our estimates are based on total cross sections since the expected number of events is so small that a distribution analysis is not possible. The distributions just helped us to find the optimum cuts. Unfortunately, competitive bounds on these anomalous couplings cannot be achieved at HERA, even with the future luminosity upgrades.\footnote{We would like to apologize for the optimistic but incorrect results that were presented at the workshop due to a regrettable and unlucky mistake in our programs.} As a counterpart, a different kinematical region is explored, in which the form factors can be neglected. \section*{Acknowledgements} One of us (J.I.) would like to thank the Workshop organizers for financial support and very especially the electroweak working group conveners and the Group from Madrid at ZEUS for hospitality and useful conversations. This work has been partially supported by the CICYT and the European Commission under contract CHRX-CT-92-0004.
proofpile-arXiv_065-424
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section*{References}
proofpile-arXiv_065-425
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{INTRODUCTION} \vskip 2mm Processes of heavy quark and lepton pair production on nucleon and nuclear targets at high energies are very interesting from both theoretical and practical reasons. These processes provide a method to study the internal structure of hadrons in the very small $x$-region. Some limits on the possible non-perturbative contributions to this region can also be obtained. Realistic estimates of the cross sections of heavy quark production are necessary in order to plan experiments on the existing and future accelerators. Predictions are usually obtained in the framework of perturbative QCD in the leading and the next-to-leading order $\alpha_s$-expansion. In the case of a nuclear target it is usually assumed that the cross section of heavy flavour production (or any other hard process), $\sigma(Q\overline{Q})$, should be proportional to $A^{\alpha}$, with $\alpha = 1$; that is, in agreement with the most accurate experimental results of [1,2], where values of $\alpha = 1.00 \pm 0.05$ for all $D$-meson production by pions at $\sqrt{s} = 22$ GeV and $\alpha = 1.02 \pm 0.03 \pm 0.02$ for neutral $D$-meson production by protons at $\sqrt{s} = 39$ GeV were obtained, respectively. However it seems important to consider the problem in more detail because, for example, many experimental results are obtained by extrapolation of results on a nuclear target to a nucleon target. It is experimentally well-known [3-9] that parton distributions in nuclei (i.e., in bound nucleons) are slightly different from the same distributions in free nucleons. So the value of $\alpha$ can differ from unity and it is interesting to estimate the size of this effect. Also the differences in the distributions of gluons, sea quarks and valence quarks are not the same \cite{Kari}. Moreover, in different hard processes different QCD subprocesses give the main contributions, so the effects on the $\alpha$-values can also be different. In this paper we calculate the values of $\alpha$ as a function of energy and Feynman-$x$ for the cases of open charm, beauty and different mass Drell-Yan pair production in proton-nucleus and nucleus-nucleus collisions. In our previous paper \cite{APSS} we showed that the results for the A-dependence of heavy flavour production depend very weakly on the used set of parton distributions. Three different sets, namely MRS-1 \cite{mrs}, MT S-DIS \cite{mt} and GRV HO \cite{grv}, which can be found in CERN PDFLIB \cite{pdf}, gave practically the same results for the $\alpha$ behaviour (see Table 1 of Ref. \cite{APSS}). Here we present the results for two sets, GRV HO \cite{grv} and MRS-A \cite{mrsa} which are in comparatively good agreement with the new HERA data and we see that they give practically identical results for both open heavy flavour and Drell-Yan pair production in proton-nucleus and nucleus-nucleus collisions. \vskip 9 mm \section{CROSS SECTIONS OF HARD PROCESSES IN QCD} \vskip 2mm The standard QCD expression for the cross section of heavy quark production in a hadron 1 - hadron 2 collision has the form \begin{equation} \sigma^{12\rightarrow Q\overline{Q}} = \int_{x_{a0}}^{1} \frac{dx_a}{x_a} \int_{x_{b0}}^{1} \frac{dx_b}{x_b}\left [x_aG_{a/1}(x_a,\mu^{2}) \right ]\left [x_bG_{b/2}(x_b,\mu^{2})\right ] \hat{\sigma}^{ab\rightarrow Q\overline{Q}}(\hat{s},m_Q,\mu^{2}) \;, \label{eq:totalpf} \end{equation} where $x_{a0} = \textstyle 4m_Q^2/ \textstyle s$ and $x_{b0} = \textstyle 4m_Q^2/ \textstyle (sx_a)$. Here $G_{a/1}(x_a,\mu^{2})$ and $G_{b/2}(x_b,\mu^{2})$ are the structure functions of partons $a$ and $b$ inside hadrons $1$ and $2$ respectively, and $\hat{\sigma}^{ab\rightarrow Q\overline{Q}}(\hat{s},m_Q,\mu^{2})$ is the cross section of the subprocess $ab\rightarrow Q\overline{Q}$ as given by standard QCD. The latter depends on the parton center-of-mass energy \mbox{$\hat{s} = (p_a+p_b)^2 = x_ax_bs$}, the mass of the produced heavy quark $m_Q$ and the QCD scale $\mu^{2}$. Eq. (\ref{eq:totalpf}) should account for all possible subprocesses $ab \rightarrow Q\overline{Q}$. The parton-parton cross section $\hat{\sigma}^{ab\rightarrow Q\overline{Q}}$ can be written in the form \cite{NDE} \begin{equation} \hat{\sigma}^{ab\rightarrow Q\overline{Q}}(\hat{s},m_Q,\mu^{2}) = \frac{\alpha^{2}_{s}(\mu^{2})}{m^{2}_{Q}}f_{ab}(\rho ,\mu^{2},m^{2}_{Q}) \;, \end{equation} with \begin{equation} \rho = 4m^{2}_{Q}/\hat{s} \end{equation} and \begin{equation} f_{ab}(\rho, \mu^2, m^2_Q) = f^{(0)}_{ab}(\rho) + 4\pi \alpha_{s}(\mu^{2})[f^{(1)}_{ab}(\rho) + \hat{f}^{(1)}_{ab}(\rho)\ln{(\mu^{2}/m^{2}_{Q})}]\;. \end{equation} The functions $f^{(0)}_{ab}(\rho)$, $f^{(1)}_{ab}(\rho)$ and $\hat{f}^{(1)}_{ab}(\rho)$ can be found in \cite{NDE} for heavy flavour production. Formulae with the same structure as (1)-(4) can be found in \cite{DY} for Drell-Yan pair production. In this case $M_{l^+l^-}$ plays the role of $2m_Q$ in $x_{a0}$, $x_{b0}$ and $\hat{\sigma}^{ab\rightarrow Q\overline{Q}}$. For the numerical calculations we wrote the nuclear structure function $G_{b/A}(x_b,\mu^{2})$ in the form \begin{equation} G_{b/A}(x_b,\mu^{2}) = A\cdot G_{b/N}(x_b,\mu^{2})\cdot R^A_b(x,\mu^{2}), \end{equation} similarly to Ref. \cite{ina} and we take the values of $R^A_b(x,\mu^{2})$ for gluons, valence and sea quarks from Ref. [10]. They are presented in Fig. 1. The values of $R^A_b(x,\mu^{2})$ in [10] are given for $x > 10^{-3}$ which is not small enough at high energies. So at $x < 10^{-3}$ we used two variants of the $R^A_b(x,\mu^{2})$ behaviour for gluon and sea quark distributions: The first is the constant frozen at $x = 10^{-3}$ (dashed-dotted curves for gluons and solid curves for sea quarks in Fig. 1). The second is the extrapolation as $x^{\beta}$ (dotted curves in Fig. 1) of the distribution which gives the main contribution to our cross section, with $\beta =$ 0.096 and 0.040 for charm and beauty production respectively (only for gluons), and $\beta$ = 0.109, 0.096 and 0.072 for Drell-Yan pair invariant masses $M^2_{l^+l^-}=$ 5, 25 and 100 GeV$^2$ respectively (only for sea quarks). Such behaviour is in qualitative agreement with the results of \cite{lev}. In the case of nucleus-nucleus collisions the parton distributions in both incident nuclei should be written in the form (5). It is necessary to note that in such a way the main part of the shadowing processes is taken into account. \vskip 9 mm \section{A-DEPENDENCE OF HEAVY FLAVOUR PRODUCTION} \vskip 2 mm In the case of charm production we have used the values $m_c$ = 1.5 GeV and $\mu^2$ = 4 GeV$^2$ ($\mu^2$ = 5 GeV$^2$ for the MRS-A set) and in the case of beauty production $m_b$ = 5 GeV and $\mu^2$ = $m_b^2$. The obtained results for $\alpha$ determined from the ratios of heavy quark production cross section on a gold target and on the proton are presented in Fig. 2 for the GRV HO and MRS-A sets\footnote{In all the calculations in this paper little difference, if any, has been found in the results for $\alpha$ between these two sets of parton distributions.} and the two variants of gluon distribution ratios in the small $x$-region. Here $\sqrt{s_{NN}}$ is the c.m. energy for the interaction of the incident proton with one target nucleon. One can see that at fixed target energies the values of $\alpha$ are slightly higher than unity, which is not in contradiction with the results of Refs. [1,2]. However $\alpha$ decreases with increasing energy and this effect is larger in the case of charm production than in the case of beauty. One can see also that the difference between the two variants for $R^A_b(x,\mu^{2})$ at $x < 10^{-3}$ becomes important only at the highest energies. We also calculate the values of $\alpha$ for different Feynman-$x$ ($x_F$) regions using $x_F = x_a - x_b$ at energies $\sqrt{s_{NN}}$ = 39 GeV and 1800 GeV. The results are presented in Fig. 3. At negative and moderate $x_F$ (in the nucleus fragmentation region) the values of $\alpha$ are slightly higher than unity. However in the case of charm production in the beam fragmentation region (positive $x_F$) the values of $\alpha$ become essentially smaller than unity. For beauty production the last effect is expected only at very high energies. The obtained results for A-dependences of charm and beauty production in the symmetric case of gold-gold collisions are presented in Figs. 4 ($\alpha$ as a function of initial energy) and 5 ($\alpha$ as a function of $x_F$ at two energies). Here $\sqrt{s_{NN}}$ is the c.m. energy for one nucleon-nucleon interaction. One can see a qualitative agreement with the case of proton-nucleus collisions (with the difference that the trivial value of $\alpha$ is here equal to two). Again we can see that the values of $\alpha$ are dependent on the initial energy and $x_F$ if the energy is high enough. Calculations for charm production in perturbative QCD (including preequilibrium charm production from secondary minijet gluons) can be found in \cite{wang} for central gold on gold collisions. \vskip 9 mm \section{ A-DEPENDENCE OF DRELL-YAN PAIR PRODUCTION} \vskip 2mm As said above, the QCD expression for heavy lepton pair production in a hadron 1 - hadron 2 collision has the same form as Eq. (1) but with another matrix element which can be found in Ref. \cite{DY}. Now we have an additional variable -- the mass of the produced pair $M_{l^+l^-}$ which plays more or less the same role as the mass of the heavy quark. However now its value can be measured experimentally and do not lead to any uncertainty. We have used the values of QCD scale $\mu^2 = M^2_{l^+l^-}$. The obtained results for $\alpha$ determined in the same way as in heavy quark production, i.e., from the ratios of Drell-Yan pair production cross section in proton-gold and proton-proton collisions, are presented in Fig. 6 for the GRV HO and MRS-A sets and the two variants of sea quark distribution ratios in the small $x$-region\footnote{As is well-known gluons are dominant for heavy flavour production, while sea quarks dominate Drell-Yan production.}. Here again $\sqrt{s_{NN}}$ is the c.m. energy for the interaction of the incident proton with one target nucleon. Contrary to the case of heavy flavour production the values of $\alpha$ are never higher than unity. They decrease more or less monotonically with increasing energy and this effect becomes smaller with increasing $l^+l^-$ mass. Again the two sets of parton distributions give practically identical results for the energy dependence of $\alpha$ and the difference between the two variants for $R^A_b(x,\mu^{2})$ at $x < 10^{-3}$ becomes important only at the highest energies. The predicted $M_{l^+l^-}$ dependence of $\alpha$ for Drell-Yan pair production in proton-nucleus collisions at different energies is also shown in Fig. 6. The results of our calculations of $\alpha$ as a function of $x_F$ for heavy lepton pair production are presented in Fig. 7. The qualitative picture is similar to the case of heavy flavour production; the numerical difference of our predictions from the value $\alpha = 1$ is here more significant for the case of a not very large mass of the lepton pair. Predictions of A-dependence for Drell-Yan pair production cross section in symmetric nucleus-nucleus collisions are shown in Figs. 8 and 9. As for the case of proton-nucleus collisions, all the effects are qualitatively the same but numerically larger than in the case of heavy flavour production. Our results are in agreement with the calculations of \cite{GGRV}. \vskip 9 mm \section{CONCLUSIONS} \vskip 2 mm In this paper we calculate the A-dependence of charm and beauty as well as Drell-Yan pair production using standard QCD formulas and accounting for the difference of parton distributions in free and bound nucleons. If one parametrize these cross sections as $\sigma \sim A^{\alpha}$, the value of $\alpha$ is slightly different from unity at the available energies. For the case of heavy flavour production at comparatively low energies the obtained values of $\alpha$ are a little higher than unity. This should be connected with some nucleon-nucleon correlations which change the large-$x$ parton distributions. In the case of Drell-Yan pair production such effect is absent because of the different relations in the contributions of valence quarks, sea quarks and gluons. At higher energies the values of $\alpha$ decrease and become smaller than unity. At $\sqrt{s_{NN}}$ = 1800 GeV we expect a value of $\alpha \sim 0.95$ for charm and low-mass Drell-Yan pair production. The decrease of the ratio $R^A_b(x,\mu^{2})$, which results in a decrease of $\alpha$, can be connected with the effects of parton density saturation \cite{glr} which in heavy nuclei occur at $x$-values higher than in the proton. If we consider two small and different values of $x_a$ and $x_b$ in Eq. (1) for the case of proton-nucleus interaction, it is clear that the contribution to the inclusive cross section from the region $x_a < x_b$ should be larger than the mirror contribution ($x_a > x_b$) because the value of the ratio $R^A_b(x,\mu^{2})$ in the first case is larger. It means that heavy quark and Drell-Yan pairs will be produced preferably in the nucleus fragmentation hemisphere, i.e., asymmetrically, which is quite usual and has been confirmed experimentally in the case of light quark production. From Figs. 3 and 7 it is clear that charm and low-mass Drell-Yan pair production on nuclear targets at LHC energies will give important information on the nuclear shadowing of the structure functions at small $x$. Besides, the experimental measurements of the effects predicted for different hard interactions should allow (in principle) to control the validity of the conventional extraction of parton distributions from the experimental DIS data. Let us repeat again that almost all nuclear shadowing corrections are accounted for in our calculations because they contribute to the $R^A_b(x,\mu^{2})$ ratios, which have been extracted from the experimental data. On the other hand these shadowing corrections are not numerically large in the considered hard processes. So we can assume that the corrections which are not accounted for do not change significantly the obtained results. In conclusion we express our gratitude to M.A.Braun for useful discussions and both to K.J.Eskola and to I.Sarcevic for sending us their numerical results. We thank the Direcci\'on General de Pol\'{\i}tica Cient\'{\i}fica and the CICYT of Spain for financial support under contract AEN96-1673. C.A.S. also thanks the Xunta de Galicia for financial support. The paper was supported in part by INTAS grant 93-0079. \newpage \noindent{\Large \bf Figure captions} \vspace{0.5cm} \noindent{\bf Fig. 1.} Functions $R^A_G(x,\mu^{2})$, $R^A_V(x,\mu^{2})$ and $R^A_S(x,\mu^{2})$, which determine the ratios of the distributions for protons in the nucleus versus free protons, for gluons (dashed-dotted and dotted curves, see text), valence quarks (dashed curves) and sea quarks (solid and dotted curves, see text) respectively, for $\mu^2 =$ 5 GeV$^2$ (upper figure) and $\mu^2 = m_b^2$ (lower figure). \noindent{\bf Fig. 2.} Energy dependence of $\alpha$ for charm and beauty production in proton-gold collisions for GRV HO (solid and dashed curves) and MRS-A (dotted and dashed-dotted curves) structure functions and using extrapolated (dashed and dashed-dotted curves) and frozen at $x= 10^{-3}$ (solid and dotted curves) ratios of gluon distributions. \noindent{\bf Fig. 3.} Feynman-$x$ dependence of $\alpha$ for charm and beauty production in proton-gold collisions at $\sqrt{s_{NN}}$ = 39 GeV (upper figure) and 1800 GeV (lower figure) for GRV HO and MSR-A structure functions and using extrapolated and frozen at $x = 10^{-3}$ ratios of gluon distributions (with the same conventions as in Fig. 2). \noindent{\bf Fig. 4.} Energy dependence of $\alpha$ for charm and beauty production in gold-gold collisions for GRV HO and MRS-A structure functions and using extrapolated and frozen at $x = 10^{-3}$ ratios of gluon distributions (with the same conventions as in Fig. 2). \noindent{\bf Fig. 5.} Feynman-$x$ dependence of $\alpha$ values for charm and beauty production in gold-gold collisions at $\sqrt{s_{NN}}$ = 39 GeV (upper figure) and 1800 GeV (lower figure) for GRV HO and MSR-A structure functions and using extrapolated and frozen at $x = 10^{-3}$ ratios of gluon distributions (with the same conventions as in Fig. 2). \noindent{\bf Fig. 6.} Mass (upper figure) and energy (lower figure) dependence of $\alpha$ for Drell-Yan pair production in proton-gold collisions for GRV HO and MRS-A structure functions and using extrapolated and frozen at $x = 10^{-3}$ ratios of sea quark distributions (with the same conventions as in Fig. 2). \noindent{\bf Fig. 7.} Feynman-$x$ dependence of $\alpha$ values for heavy lepton pair production in proton-gold collisions at $\sqrt{s_{NN}}$ = 39 GeV (upper curves in each figure) and 1800 GeV (lower curves in each figure) and different masses for GRV HO and MSR-A structure functions and using extrapolated and frozen at $x = 10^{-3}$ ratios of sea quark distributions (with the same conventions as in Fig. 2). Note that at $\sqrt{s_{NN}}$ = 39 GeV all curves coincide. \noindent{\bf Fig. 8.} Mass (upper figure) and energy (lower figure) dependence of $\alpha$ for Drell-Yan pair production in gold-gold collisions for GRV HO and MRS-A structure functions and using extrapolated and frozen at $x = 10^{-3}$ ratios of sea quark distributions (with the same conventions as in Fig. 2). \noindent{\bf Fig. 9.} Feynman-$x$ dependence of $\alpha$ values for heavy lepton pair production in gold-gold collisions at $\sqrt{s_{NN}}$ = 39 GeV (upper curves in each figure) and 1800 GeV (lower curves in each figure) and different masses for GRV HO and MSR-A structure functions and using extrapolated and frozen at $x = 10^{-3}$ ratios of sea quark distributions (with the same conventions as in Fig. 2). Note that at $\sqrt{s_{NN}}$ = 39 GeV all curves coincide. \newpage
proofpile-arXiv_065-426
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} In the last decades there has been an intensive activity in studying (super)particles and (super) strings by use of different approaches aimed at finding a formulation, which would be the most appropriate for performing the covariant quantization of the models. Almost all of the approaches use twistor variables in one form or another \cite{penrose} -- \cite{bpstv}. This allowed one to better understand the geometrical and group--theoretical structure of the theory and to carry out a covariant Hamiltonian analysis (and in some cases even the covariant quantization) of (super)particle and (super)string dynamics in space--time dimensions $D=3,4,6$ and 10, where conventional twistor relations take place. It has been shown that twistor--like variables appear in a natural way as superpartners of Grassmann spinor coordinates in a doubly supersymmetric formulation \cite{spinsup} of Casalbuoni--Brink--Schwarz superparticles and Green--Schwarz superstrings \cite{gsw}, the notorious fermionic $\kappa$--symmetry \cite{ks} of these models being replaced by more fundamental local supersymmetry on the worldsheet supersurface swept by the superparticles and superstrings in target superspace \cite{stvz}. This has solved the problem of infinite reducibility of the fermionic constraints associated with $\kappa$--symmetry \footnote{A comprehensive list of references on the subject the reader may find in \cite{bpstv}}. As a result new formulation and methods of quantization of $D=4$ compactifications of superstrings with manifest target--space supersymmetry have been developed (see \cite{ber} for a review). However, the complete and simple solution of the problem of $SO(1,D-1)$ covariant quantization of twistor--like superparticles and superstrings in $D~>~4$ is still lacking. To advance in solving this problem one has to learn more on how to deal with twistor--like variables when performing the Hamiltonian analysis and the quantization of the models. In this respect a bosonic relativistic particle in a twistor--like formulation may serve as the simplest but rather nontrivial toy model. The covariant quantization of the bosonic particle has been under intensive study with both the operator and path--integral method \cite{ferber,teit,polyakov,govaerts,sf,sg,bh,bfortschr}. In the twistor--like approach the bosonic particle has been mainly quantized by use of the operator formalism. For that different but classically equivalent twistor--like particle actions have been considered \cite{ferber,es,sg,bh,bfortschr}. The aim of the present paper is to study some features of bosonic particle path--integral quantization in the twistor--like approach by use of the BRST--BFV quantization prescription \cite{bf} -- \cite{bff}. In the course of the Hamiltonian analysis we shall observe links between various formulations of the twistor--like particle \cite{ferber,es,stvz} by performing a conversion of the Hamiltonian constraints of one formulation to another. A particular feature of the conversion procedure \cite{fs} applied to turn the second--class constraints into the first--class constraints is that the simplest Lorentz--covariant way to do this is to convert a full mixed set of the initial first-- and second--class constraints rather than explicitly extracting and converting only the second--class constraints. Another novel feature of the conversion procedure applied below (in comparison with the conventional one \cite{bff,fs}) is that in the case of the $D=4$ and $D=6$ twistor--like particle the number of new auxiliary Lorentz--covariant coordinates, which one introduces to get a system of first--class constraints in an extended phase space, exceeds the number of independent second--class constraints of the original dynamical system, (but because of an appropriate amount of the first--class constraints we finally get, the number of physical degrees of freedom remains the same). In Section 2 the classical mechanics of a twistor--like bosonic particle in D=3,4 and 6 is considered. The Hamiltonian analysis of the constraints accompanied by the conversion procedure is carried out and a classical BRST charge is constructed by introducing ghosts corresponding to a set of the first--class constraints obtained as a result of conversion. In Section 3 the problem of admissible gauge choice for variables describing the matter--ghost system of the model is discussed. In Section 4 we perform the path--integral quantization of the model in $D=3,4$ and $6$ space--time dimensions using the extended BRST scheme \cite{mhenn}. We calculate the particle propagator and show, that it coincides with that of the massless bosonic particle. At the end of this Section we make a comment on problems of the D=10 case. {\it Notation.} We use the following signature for the space-time metrics: $(+,-,...,-)$. \section{Classical Hamiltonian dynamics and the BRST-charge.} \subsection{Preliminaries} The dynamics of a massless bosonic particle in D=3,4,6 and 10 space--time can be described by the action \cite{ferber} \begin{equation} \label{201} S={1\over 2}\int d\tau \dot x^m ({\bar \l}\gamma _m \l ), \qquad \end{equation} where $x^m(\tau )$ is a particle space--time coordinate, $\l ^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}(\tau )$ is an auxiliary bosonic spinor variable, the dot stands for the time derivative $\partial\over{\partial\tau}$ and $\gamma ^m $ are the Dirac matrices. The derivation of the canonical momenta \footnote{In what follows $P^{(..)}$ denotes the momentum conjugate to the variable in the brackets} $P^{(x)}_m={{\partial L} \over{\partial{\dot x}^m}},~~ P^{(\l )}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}={{\partial L}\over {\partial}{{\dot \l}^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}}}$ results in a set of primary constraints $$ \Psi _m=P^{(x)}_m-{1\over 2}({\bar \l}\gamma _m \l )\approx 0, $$ \begin{equation} \label{202} P^{(\l )}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}\approx 0. \qquad \end{equation} They form the following algebra with respect to the Poisson brackets\footnote{The canonical Poisson brackets are $$ [P^{(x)}_m, x^n]_P=\delta ^n_m; \qquad [P^{(\l )}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}, \l^{\b}]_P=\delta ^{\b}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta} $$} \begin{equation} \label{203} [\Psi _m, \Psi _n]_P=0,~~ [P^{(\l )}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}, P^{(\l )}_{\b}]_P=0,~~ [\Psi _m, P^{(\l )}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}]_P=({\gamma _m}\l )_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}. \qquad \end{equation} One can check that new independent secondary constraints do not appear in the model. In general, Eqs. \p{202} are a mixture of first-- and second--class constraints. The operator quantization of this dynamical system in $D=4$ (considered previously in \cite{sg,bh}) was based on the Lorentz--covariant splitting of the first-- and second--class constraints and on the subsequent reduction of the phase space (either by explicit solution of the second--class constraints \cite{bh} or, implicitly, by use of the Dirac brackets \cite{sg}), while in \cite{bz0,bfortschr} a conversion prescription \cite{bff,fs} was used. The latter consists in the extension of the phase space of the particle coordinates and momenta with auxiliary variables in such a way, that new first--class constraints replace the original second--class ones. Then the initial system with the second--class constraints is treated as a gauge fixing of a ``virtual" \cite{bff} gauge symmetry generated by the additional first--class constraints of the extended system \cite{bff,fs}. This is achieved by taking the auxiliary conversion degrees of freedom to be zero or expressed in terms of initial variables of the model. The direct application of this procedure can encounter some technical problems for systems, where the first-- and second--class constraints form a complicated algebra (see, for example, constraints of the $D=10$ superstring in a Lorentz--harmonic formulation \cite{bzstr}). Moreover, in order to perform the covariant separation of the first-- and second--class constraints in the system under consideration it is necessary either to introduce one more independent auxiliary bosonic spinor $\mu _{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}$ (the second component of a twistor $Z^A=(\l^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta},\mu _{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta})$ \cite{penrose}) or to construct the second twistor component from the variables at hand by use of a Penrose relation \cite{penrose} ${\bar \mu}^{\dot \alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}=ix^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta {\dot \alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}}\l _{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta} ~(D=4),~ \mu^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}=x^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta \b}\l _{\b} ~(D=3)$. In the latter case the structure of the algebra of the first-- and second--class constraints separated this way \cite{sfortschr,sg} makes the conversion procedure rather cumbersome. To elude this one can try to simplify the procedure by converting into the first class the whole set \p{202} of the mixed constraints. The analogous trick was used to convert fermionic constraints in superparticle models \cite{es,moshe}. Upon carrying out the conversion procedure we get a system characterized by the set of first--class constraints $T_i$ that form (at least on the mass shell) a closed algebra with respect to the Poisson brackets defined for all the variables of the modified phase space. In order to perform the BRST--BFV quantization procedure we associate with each constraint of Grassmann parity $\epsilon $ the pair of canonical conjugate auxiliary variables (ghosts) $\eta _i,~ P^{(\eta )}_i$ with Grassmann parity $\epsilon +1$ \footnote{If the extended BRST--BFV method is used, with each constraint associated are also a Lagrange multiplier, its conjugate momentum of Grassmann parity $\epsilon $ and an antighost and its momentum of Grassmann parity $\epsilon +1$ (see \cite{bf,mhenn} for details).}. The resulting system is required to be invariant under gauge transformations generated by a nilpotent fermionic BRST charge $\Omega $. This invariance substitutes the gauge symmetry, generated by the first class constraints in the initial phase space. The generator $\Omega $ is found as a series in powers of ghosts $$ \Omega ={\eta _i}T_i + higher~ order~ terms, $$ where the structure of higher--order terms reflects the noncommutative algebraic structure of the constraint algebra \cite{mhenn}. Being the generator of the BRST symmetry $\Omega$ must be a dynamical invariant: $$ {\dot \Omega}=[\Omega ,H]_P=0, $$ where $H$ is a total Hamiltonian of the system, which has the form \begin{equation}\label{h} H=H_0+[\chi ,\Omega]_P. \end{equation} In \p{h} $H_0$ is the initial Hamiltonian of the model and $\chi$ is a gauge fixing fermionic function whose form is determined by admissible gauge choices \cite{teit,govaerts,sf,west,bfortschr} (see Section 3 for the discussion of this point). Upon quantization $\Omega$ and $H$ become operators acting on quantum state vectors. The physical sector of the model is singled out by the requirement that the physical states are BRST invariant and vanish under the action of $\Omega$. Another words, we deal with a {\it quantum gauge theory}. When the gauge is fixed, we remain only with physically nonequivalent states, and the Hamiltonian $H$ is argued to reproduce the correct physical spectrum of the quantum theory. When the model is quantized by the path--integral method, we also deal with a quantum gauge theory. The Hamiltonian \p{h} is used to construct an effective action and a corresponding BRST-invariant generating functional which allows one to get transition amplitudes between physical states of the theory. Below we consider the conversion procedure and construct the BRST charge for the twistor--like particle model in dimensions $D=3,4$ and $6$. \subsection{ D=3} In $D=3$ the action \p{201} is rewritten as \begin{equation}\label{211} S={1\over 2}\int d\tau \l ^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}{\dot x}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta \b} \l ^{\b}, \qquad \end{equation} where $\l^\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta$ is a real two-component commuting spinor (spinor indices are risen and lowered by the unit antisymmetric tensor $\epsilon _{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta \b}$) and $x_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta \b}=x_m\gamma ^m_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta \b}$. The system of primary constraints \p{202} \begin{eqnarray} \label{212} \Psi_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta \b}=P^{(x)}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta \b}-\l_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}\l_{\b}\approx 0, \nonumber \\ P^{(\l )}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}\approx 0, \qquad \end{eqnarray} is a mixture of a first--class constraint generating the $\tau$--reparametrization transformations of $x$ $$ \phi = \l^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}P^{(x)}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta \b}\l^{\b} $$ and four second--class constraints \begin{equation} \label{2125} (\l P^{(\l )}),~~ (\mu P^{(x)}\mu )-(\l \mu )^{2}, ~~(\mu P^{(\l )}),~~(\l P^{(x)} \mu ), \qquad \end{equation} where $\mu ^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}=x^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta \b}\l _{\b}$ (see \cite{sg} for details). In order to perform a conversion of \p{212} into a system of first--class constraints we introduce a pair of canonical conjugate bosonic spinors $(\z^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta},~P^{(\z )}_{\b})$, $[P^{(\z )}_{\b}, \z^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}]_P=\delta ^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta} _{\b}$, and take the modified system of constraints, which is of the first class, in the following form: \begin{eqnarray} \label{213} \Psi '_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta \b}=P^{(x)}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta \b}-(\l_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}-\z_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta})(\l_{\b}-\z_{\b})\approx 0, \nonumber \\ \Phi '_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}=P^{(\l )}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}+P^{(\z )}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}\approx 0. \end{eqnarray} Eqs. \p{213} reduce to \p{212} by putting the auxiliary variables $\z^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}$ and $P^{(\z )}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}$ equal to zero. This reflects the appearance in the model of a new gauge symmetry with respect to which $\z^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}$ and $P^{(\z )}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}$ are pure gauge degrees of freedom. It is convenient to choose the following phase--space variables as independent ones: \begin{eqnarray} \label{214} v^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}=\l^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}-\z^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}, \qquad P^{(v)}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}={1\over 2}(P^{(\l )}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}-P^{(\z )} _{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}), \nonumber \\ w^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}=\l^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}+\z^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}, \qquad P^{(w)}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}={1\over 2}(P^{(\l )}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}+P^{(\z )}_ {\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}), \qquad \end{eqnarray} Then Eqs. \p{213} take the following form \begin{eqnarray} \label{215} \Psi '_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta \b}=P^{(x)}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta \b}-v_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}v_{\b}\approx0, \nonumber \\ P^{(w)}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}\approx 0. \qquad \end{eqnarray} These constraints form an Abelian algebra. One can see that $w^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}$ variables do not enter the constraint relations, and their conjugate momenta are zero. Hence, the quantum physical states of the model will not depend on $w^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}$ . Enlarging the modified phase space with ghosts, antighosts and Lagrange multipliers in accordance with the following table \begin{tabular}{cccc} &&&\\ Constraint & Ghost & Antighost & Lagrange~multiplier \\ ${\Psi}'_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta \b}$ & $c^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta \b}$ & $\tilde c^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta \b}$ & $e^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta \b}$ \\ $P^{(w)}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}$ & $b^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}$ & $\tilde b^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}$ & $f^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}$ \\ &&& \end{tabular} \noindent we write the classical BRST charges \cite{bf,mhenn} of the model in the minimal and extended BRST--BFV version as follows \begin{equation} \label{218} \Omega_{min}=c^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta \b}\Psi '_{\b \alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}+b^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}P^{(w)}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}, \qquad \end{equation} \begin{equation} \label{217} \Omega =P^{(\tilde c)}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta \b}P^{(e)\b \alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}+P^{(\tilde b) \alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}P^{(f)}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}+ \Omega _{min}. \qquad \end{equation} \subsection{D=4} In this dimension we use two--component $SL(2,C)$ spinors $(\l^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}=\epsilon^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta \b}\l_{\b}; ~{\bar \l}^{\dot \alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}=\epsilon^{{\dot \alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta} {\dot \b}}{\bar \l}_{\dot \b}; ~\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta ,{\dot \alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}=1,2; ~ \epsilon^{12}=-\epsilon_{21}=1)$. Other notation coincides with that of the $D=3$ case. Then in $D=4$ the action \p{201} can be written as following \begin{equation}\label{221} S= {1\over 2}\int d \tau \l^\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta {\dot x}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta {\dot \alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}}\bar{\l^{\dot \alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}}, \qquad \end{equation} where $ x_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta\dot \alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}= x_{m}{\sigma}^{m}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta {\dot \alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}} $, and ${\sigma}^m_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta {\dot \alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}}$ are the relativistic Pauli matrices. The set of the primary constraints \p{202} in this dimension $$\Psi_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta {\dot {\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}}}= P^{(x)}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta {\dot {\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}}}- \bar{ \l}_{\dot \alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta} \l_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta} \approx 0,$$ \begin{equation} \label{222} { P^{(\l )}}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}\approx 0, \qquad \end{equation} $${\bar P}^{(\bar {\l} )}_{\dot {\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}}\approx 0 $$ contains two first--class constraints and three pairs of conjugate second--class constraints \cite{sg,sfortschr}. One of the first class constraints generates the $\tau$-reparametrization transformations of $x^{{\dot\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta} \alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}$ $$ \phi= \l^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}P^{(x)}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta {\dot {\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}}}{\bar {\l}}^{\dot {\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}} $$ and another one generates $U(1)$ rotations of the complex spinor variables \begin{equation} \label{2221} U=i(\l^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta} P^{(\l )}_\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta- \bar {\l}^{\dot{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}}\bar P^{(\l )}_{\dot {\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}}). \end{equation} The form of the second--class constraints is analogous to that in the D=3 case (see Eq. \p{2125} and \cite{sg}), and we do not present it explicitly since it is not used below. To convert the mixed system of the constraints \p{222} into first--class constraints one should introduce at least three pairs of canonical conjugate auxiliary bosonic variables, their number is to be equal to the number of the second--class constraints in \p{222}. However, since we do not want to violate the manifest Lorentz invariance, and the $D=4$ Lorentz group does not have three--dimensional representations, we are to find a way round. We introduce two pairs of canonical conjugate conversion spinors $ (\zeta^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta},P^{(\zeta )}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}),~~ [\zeta^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta},P^{(\zeta)}_{\b}]_P=-\delta^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}_{\b}, ~~[\bar {\zeta}^{\dot {\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}},{\bar P}^{(\bar \zeta )}_{\dot {\b}}]_P=-\delta^ {\dot \alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}_{\dot \b}, $ (i.e. four pairs of real auxiliary variables) and modify the constraints \p{222} and the $U(1)$ generator, which becomes an independent first--class constraint in the enlarged phase space. Thus we get the following system of the first--class constraints: $$\Psi^{'}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta \dot {\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}}=P^{(x)}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta \dot {\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}} -(\bar {\l}-\bar {\zeta})_{\dot {\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}} (\l-\zeta)_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}\approx 0,$$ \begin{equation} \label{223} { \Phi_\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}=P^{(\l )}_\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta+P^{(\zeta )}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}\approx 0 , \qquad \end{equation} $$ \bar {\Phi}_{\dot {\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}}= \bar P^{(\bar \l )}_{\dot {\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}}+\bar P^{(\bar \zeta )}_{\dot {\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}}\approx 0 , $$ $$ U=i(\l^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta} P^{(\l )}_\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta+\zeta^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta} P^{(\zeta )}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}- \bar {\l}^{\dot {\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}}\bar P^{(\l )}_{\dot {\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}}-\bar {\zeta}^{\dot {\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}}\bar P^{(\zeta )}_{\dot {\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}})\approx 0. $$ One can see (by direct counting), that the number of independent physical degrees of freedom of the particle in the enlarged phase space is the same as in the initial one. The latter is recovered by imposing gauge fixing conditions on the new auxiliary variables \begin{equation} \label{2231} \z^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}=0,\qquad{\bar \z}^{\dot \alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}=0,\qquad P^{(\z )}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}=0, \qquad P^{({\bar \z})}_{\dot \alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}=0. \end{equation} By introducing a new set of the independent spinor variables analogous to that in \p{214} one rewrites Eqs. \p{223} as follows $$ \Psi'_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta {\dot{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}}}=P^{(x)}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta {\dot{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}}}- v_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}{\bar v}_{\dot{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}}\approx 0, $$ $$ U=i(P^{(v)}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}v^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}- P^{(\bar v)}_{\dot {\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}}{\bar v}^{\dot {\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}})\approx 0, $$ \begin{equation} \label{224} { P^{(w)}}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}\approx 0, \qquad \end{equation} $$ P^{(\bar{w})}_{\dot{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}}\approx 0. $$ Again, as in the $D=3$ case, $w_\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta,~{\bar w}_{\dot\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}$ and their momenta decouple from the first pair of the constraints \p{224}, and can be completely excluded from the number of the dynamical degrees of freedom by putting \begin{equation}\label{es} w_\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta=\l_\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta+\z_\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta=0, \qquad {P^{(w)}}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}={1\over 2}({ P^{(\l )}}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}+{P^{(\z )}}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta})=0 \end{equation} in the strong sense. This gauge choice, which differs from \p{2231}, reduces the phase space of the model to that of a version of the twistor--like particle dynamics, subject to the first pair of the first--class constraints in \p{224}, considered by Eisenberg and Solomon \cite{es}. The constraints \p{224} form an abelian algebra, as in the $D=3$ case. In compliance with the BRST--BFV prescription we introduce ghosts, antighosts and Lagrange multipliers associated with the constraints \p{224} as follows \begin{tabular}{cccc} Constraint & Ghost & Antighost & Lagrange~ multiplier \\ ${\Psi} '_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta {\dot{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}}}$ & $c^{\dot{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta} \alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}$ & ${\tilde c}^{{\dot \alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta} \alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}$ & $e^{\dot{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta} \alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta} $ \\ $U$ & $a$ & ${\tilde a}$ & $g$ \\ $P^{(w)}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}$ & $b^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}$ & ${\tilde b}^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}$ & $f^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}$ \\ $P^{(\bar{w})}_{\dot{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}}$ & ${\bar b}^{\dot{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}}$ & ${\tilde{\bar b}}^{\dot{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}}$ & ${\bar f}^{\dot {\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}}$ \\ \qquad \end{tabular} Then the BRST--charges of the $D=4$ model have the form \begin{equation} \label{227} \Omega_{min}=c^{\dot {\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta} \alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}\Psi_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta {\dot {\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}}}+b^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}P^{(w)}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}+ {\bar b}^{\dot{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}}P^{({\bar w})}_{\dot{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}}+aU, \qquad \end{equation} \begin{equation} \label{226} \Omega=P^{({\tilde c})}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta {\dot{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}}}P^{(e) {\dot{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}}\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}+P^{({\tilde b})} _{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}P^{(f)\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}+P^{({\tilde{\bar b}})}_{\dot{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}}P^{({\bar f})\dot{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}}+ P^{({\tilde a})}P^{(g)}+\Omega_{min}. \end{equation} \subsection{D=6} In $D=6$ a light--like vector $V^m$ can be represented in terms of commuting spinors as follows $$ V^m=\l^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}_i\gamma^m_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta \b}\l^{\b i}, $$ where $\l^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}_i$ is an $SU(2)$--Majorana--Weyl spinor which has the $SU^*(4)$ index $\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta =1,2,3,4$ and the $SU(2)$ index $i=1,2$. $\gamma^m_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta \b}$ are $D=6$ analogs of the Pauli matrices (see \cite{6spin,benght}). $SU(2)$ indices are risen and lowered by the unit antisymmetric tensors $\epsilon_{ij},~~\epsilon^{ij}$. As to the $SU^*(4)$ indices, they can be risen and lowered only in pairs by the totally antisymmetric tensors $\epsilon_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta \b \g \delta}\def\Om{\Omega}\def\s{\sigma}$, $\epsilon^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta \b \g \delta}\def\Om{\Omega}\def\s{\sigma}$ ($\epsilon_{1234}=1)$. Rewriting the action \p{201} in terms of $SU(2)$--Majorana--Weyl spinors, one gets \begin{equation} \label{231} S={1\over 2}\int d \tau {\dot x}^m \l^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}_i(\gamma_m)_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta \b}\l^{\b i}. \end{equation} The system of the primary constraints \p{202} takes the form $$ \Psi_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta \b}=P^{(x)}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta \b}-\epsilon_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta \b \g \delta}\def\Om{\Omega}\def\s{\sigma}\l^{\g}_i \l^{\delta}\def\Om{\Omega}\def\s{\sigma i} \approx 0, $$ \begin{equation} \label{232} { ~P^{(\l )i}}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}\approx 0, \qquad \end{equation} where $P^{(x)}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta \b}=P^{(x)}_m \gamma^m_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta \b}$. $\Psi_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta \b}$ is antisymmetric in $\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta$ and $\b$ and contains six independent components. (To get \p{232} we used the relation $(\g_m)_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta\b}\g^m_{\g\delta}\def\Om{\Omega}\def\s{\sigma}\sim \epsilon_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta \b \g \delta}\def\Om{\Omega}\def\s{\sigma}$). From Eqs. \p{232} one can separate four first--class constraints by projecting \p{232} onto $\l^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}_{i}$ \cite{sfortschr,benght}. One of the first--class constraints generates the $\tau$--reparametrizations of $x^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta\b}$ $$ \phi =\l^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}_{i}P^{(x)}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta \b}\l^{\b i}, $$ and another three ones form an $SU(2)$ algebra $$ T_{ij}=\l^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}_{(i}P^{(\l )}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta j)}, $$ Braces denote the symmetrization of $i$ and $j$. All other constraints in \p{232} are of the second class. The conversion of \p{232} into first--class constraints is carried out by analogy with the $D=4$ case. According to the conventional conversion prescription we had to introduce five pairs of canonical conjugate bosonic variables. Instead, in order to preserve Lorentz invariance, we introduce the canonical conjugate pair of bosonic spinors $\z^{\b}_j$, $P^{(\z )i}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}$ ($[P^{(\z )i}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta},\z^{\b}_j]_P=\delta}\def\Om{\Omega}\def\s{\sigma^{\b}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}\delta}\def\Om{\Omega}\def\s{\sigma^i_j,$) modify the constraints \p{232} and the $SU(2)$ generators. This results in the set of independent first--class constraints $$ \Psi '_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta \b}=P^{(x)}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta \b}-\epsilon_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta \b \g \delta}\def\Om{\Omega}\def\s{\sigma} (\l^{\g}_i-\z^{\g}_i)(\l^{\delta}\def\Om{\Omega}\def\s{\sigma i}-\z^{\delta}\def\Om{\Omega}\def\s{\sigma i})\approx 0, $$ \begin{equation} \label{233} { \Phi_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}}^i=P^{(\l )i}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}+P^{(\z )i}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}\approx 0, \qquad \end{equation} $$ T_{ij}=\l^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}_{(i}P^{(\l )}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta j)}-\z^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}_{(i}P^{(\z )}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta j)}\approx 0. $$ In terms of spinors $v^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}_{i}$ and $w^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}_{i}$, and their momenta, defined as in the $D=3$ case \p{214}, they take the following form $$ \Psi '_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta \b}=P^{(x)}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta \b}-\epsilon_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta \b \g \delta}\def\Om{\Omega}\def\s{\sigma}v^{\g}_iv^{\delta}\def\Om{\Omega}\def\s{\sigma i}\approx 0, $$ \begin{equation} \label{234} { T_{ij}}=v^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}_{(i}P^{(v)}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta j)}\approx 0, \qquad \end{equation} $$ P^{(w)i}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}\approx 0. $$ These constraints form a closed algebra with respect to the Poisson brackets. The only nontrivial brackets in this algebra are \begin{equation} \label{235} [T_{ij},T_{kl}]_{p}=\epsilon_{jk}T_{il}+\epsilon_{il}T_{jk}+ \epsilon_{ik}T_{jl} +\epsilon_{jl}T_{ik}, \qquad \end{equation} which generate the $SU(2)$ algebra. We introduce ghosts, antighosts and Lagrange multipliers related to the constraints \p{235} \begin{tabular}{cccc} & & &\\ Constraint & Ghost & Antighost & Lagrange~ multiplier \\ ${\Psi} '_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta \b}$ & $c^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta \b}$ & ${\tilde c}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta \b}$ & $e^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta \b}$ \\ $T_{ij}$ & $a^{ij}$ & ${\tilde a}_{ij}$ & $g^{ij}$ \\ $\Phi ^{i}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}$ & $b^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}_{i}$ & ${\tilde b}^{i}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}$ & $f^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}_{i}$ \\ &&& \end{tabular} \noindent and construct the BRST charges corresponding respectively, to the minimal and extended BRST--BFV version, as follows \begin{equation} \label{238} \Omega_{min}=c^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta \b}\Psi '_{\b \alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}+b^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}_{i}P^{(w)i}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}+a^{ij}T_{ji}+ \qquad \end{equation} $$ (\epsilon_{jk}P^{(a)}_{il}+ \epsilon_{il}P^{(a)}_{jk}+ \epsilon_{ik}P^{(a)}_{jl}+\epsilon_{jl}P^{(a)}_{ik})a^{ij}a^{kl}. $$ \begin{equation} \label{237} \Omega=P^{({\tilde c})}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta \b}P^{(e)\b \alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}+P^{({\tilde b})\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta} _{i}P^{(f)i}_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}+P^{({\tilde a})ij}P^{(g)}_{ji}+\Omega_{min}, \qquad \end{equation} Higher order terms in ghost powers appear in \p{238} and \p{237} owing to the noncommutative $SU(2)$ algebra of the $T_{ij}$ constraints \p{235}. \section{Admissible gauge choice.} One of the important problems in the quantization of gauge systems is a correct gauge choice. In the frame of the BRST--BFV quantization scheme gauge fixing is made by an appropriate choice of the gauge fermion that determines the structure of the quantum Hamiltonian. The Batalin and Vilkovisky theorem \cite{bf,mhenn} reads that the result of path integration does not depend on the choice of the gauge fermions if they belong to the same equivalence class with respect to the BRST--transformations. An analogous theorem takes place in the operator BRST--BFV quantization scheme \cite{sf}. Further analysis of this problem for systems possessing the reparametrization invariance showed that the result of path integration does not depend on the choice of the gauge fermion if only appropriate gauge conditions are compatible with the boundary conditions for the parameters of the corresponding gauge transformations \cite{polyakov,govaerts,sf,west,bfortschr}. In particular, it was shown that the so--called ``canonical gauge", when the worldline gauge field of the reparametrization symmetry of the bosonic particle is fixed to be a constant, is not admissible in this sense. (see \cite{govaerts,bfortschr} for details). Anyway one can use the canonical gauge as a consistent limit of an admissible gauge \cite{sf}. Making the analysis of the twistor--like model one can show that admissible are the following gauge conditions on Lagrange multipliers from the corresponding Tables of the previous section in the dimensions $D=3,$ 4 and $6$ of space--time, respectively, \begin{equation} \label{261} D=3:\qquad {\dot e}^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta\b}=0;\qquad f^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}=0; \end{equation} \begin{equation} \label{262} D=4:\qquad {\dot e}^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta\dot\b}=0;\qquad f^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}=0; \qquad f^{\dot \alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}=0;\qquad g=0; \end{equation} \begin{equation} \label{263} D=6:\qquad {\dot e}^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta\b}=0;\qquad f_i^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}=0;\qquad g^{ij}=0; \end{equation} The canonical gauge \begin{equation} \label{264} e=constant, \end{equation} can be considered as a limit of more general admissible gauge $e-\varepsilon {\dot e}=constant$ (at $\varepsilon \rightarrow 0$) \cite{sf}. Then the use of the gauge condition \p{264} does not lead to any problems with the operator BRST--BFV quantization. Below we shall use the ``relativistic" gauge conditions \p{261}, \p{262} and \p{263} for the path--integral quantization. The use of the canonical gauge \p{264} in this case would lead to a wrong form of the particle propagator. \section{Path--integral BRST quantization.} In this section we shall use the extended version of the BRST--BFV quantization procedure \cite{mhenn,bff} and fix the gauge by applying the conditions \p{261}, \p{262}, \p{263}. The gauge fermion, corresponding to this gauge choice, is \begin{equation} \label{417} \chi_D={1\over 2}P^{(c)}_me^m, ~~~D=3,4,6, \qquad \end{equation} The Hamiltonians constructed with \p{417} are \cite{bf,mhenn} $$ {\it H}_D=[\Omega_D,\chi_D],\qquad D=3,4,6 $$ \begin{equation} \label{420} {H}_3=e^{m}(P^{(x)}_{m}-{1\over 2}v^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}(\gamma_{m})_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta \b}v^{\b}) -P^{(c)}_{m}P^{({\tilde c})m}, \qquad \end{equation} \begin{equation} \label{421} {H}_4=e^{m}(P^{(x)}_{m}-{1\over 2}{\bar v}^{\dot {\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}}(\sigma_{m}) _{\dot{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}v^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta})-P^{(c)}_{m}P^{({\tilde c})m}, \qquad \end{equation} \begin{equation} \label{422} {H}_6=e^{m}(P^{(x)}_{m}-{1\over 2}v^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}_i(\g_{m})_ {\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta \b}v^{\b i}) -P^{(c)}_{m}P^{({\tilde c})m}, \qquad \end{equation} We shall calculate the coordinate propagator $Z=\langle x_{1}^{m}\mid U_0\mid x_{2}^{m}\rangle$ (where $U_0=expiH(T_1-T_2)$ is the evolution operator), therefore boundary conditions for the phase space variables are fixed as follows: \begin{equation} \label{423} x^{m}(T_1)=x_{1}^{m}, \qquad x^{m}(T_2)=x_{2}^{m}, \qquad \end{equation} the boundary values of the ghosts, antighosts and canonical momenta of the Lagrange multipliers are put equal to zero (which is required by the BRST invariance of the boundary conditions \cite{mhenn}), and we sum up over all possible values of the particle momentum and the twistor variables. The standard expression for the matrix element of the evolution operator is \begin{equation} \label{424} Z_D=\int[D\mu DP^{\mu}]_D exp(i\int_{T_1}^{T_2} d \tau ([P^{\mu}{\dot {\mu}}]_D -{\it H}_D )), \qquad D=3,4,6. \end{equation} $[D\mu DP^{\mu}]_D $ contains functional Liouville measures of all the canonical variables of the BFV extended phase space \cite{bf}. $[P^{\mu}{\dot{\mu}}]_D $ contains a sum of products of the canonical momenta with the velocities. For instance, an explicit expression for the path--integral measure in the $D=3$ case is $$ [D\mu DP^{\mu}]=DxDP^{(x)}DvDP^{(v)}DwDP^{(w)}DeDP^{(e)}DfDP^{(f)} $$ $$ DbDP^{(b)}DcDP^{(c)}D{\tilde b}DP^{({\tilde b})}D{\tilde c}DP^{({\tilde c})}. $$ We can perform straightforward integration over the all variables that are not present in the Hamiltonians \p{420}, \p{421}, \p{422} \footnote{ All calculations are done up to a multiplication constant, which can always be absorbed by the integration measure.}. Then \p{424} reduces to the product of two terms \begin{equation} \label{425} Z_D=I_DG_D, \qquad \end{equation} where \begin{equation} \label{426} G_D=\int DcDP^{(c)}D{\tilde c}DP^{({\tilde c})} exp(i\int_{T_1}^{T_2} d \tau (P^{(c)}_m{\dot c}^m+P^{({\tilde c})}_m{\dot{\tilde c}}^m -{1\over 2} P^{({\tilde c})}_mP^{(c)m})), \qquad \end{equation} and $I_D$ includes the integrals over bosonic variables entering \p{420}, \p{421}, \p{422} together with their conjugated momenta. We use the method analogous to that in \cite{rivelles} for computing these integrals. The calculation of the ghost integral $G_D$ results in \begin{equation} \label{427} G_D=(\Delta T)^D, \qquad \Delta T=T_2-T_1, \qquad D=3,4,6. \end{equation} Let us demonstrate main steps of the $I_D$ calculation in the $D=3$ case \begin{eqnarray}\label{801} I_3&=&\int DxDP^{(x)}DeDP^{(e)}DvDP^{(v)}exp(i\int _{T_1}^{T_2}d\tau (P_m^{(x)}{\dot x}^m+P_m^{(e)}{\dot e}^m+ P_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}^{(v)}{\dot v}^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta} \nonumber \\ & & -e^{m}(P^{(x)}_{m}- {1\over 2}v^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}(\g _m)_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta \b} v^{\b})) \end{eqnarray} Integration over $P_m^{(e)}$ and $P_m^{(v)}$ results in the functional $\delta$-functions $\delta ({\dot e}),~ \delta ({\dot v})$ which reduce functional integrals over $e^m$ and $v^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}$ to ordinary ones: \begin{equation} \label{802} I_3=\int DxDP^{(x)}d^{3}ed^{2}v~exp(ip_m\Delta x^m- \\ i\int _{T_1}^{T_2}d\tau (x^m{\dot P}^{(x)}_m+e^m(P^{(x)}_m- {1\over 2}v^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}(\g _m)_{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta \b}v^{\b})), \qquad \end{equation} where $\Delta x^m=x_2^m-x_1^m$ \p{423}. Since the integral over $v^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}$ is a usual Gauss integral after integrating over $x^m$ and $v^{\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta}$ one obtains \begin{equation} \label{803} I_3=\int d^{3}pd^{3}e{1\over {\sqrt{e^me_m-i0}}}exp(i(p_m\Delta x^m- e^mp_m\Delta T)). \end{equation} In general case of $D=3$, 4 and 6 dimensions, one obtains \begin{equation} \label{428} I_D=\int d^Dpd^De{{1\over{(e^me_m-i0)^{{D-2}\over 2}}}} exp(i(p_m\Delta x^m- e^mp_m\Delta T)), \end{equation} that can be rewritten as \begin{equation} \label{804} I_D=\int d^{D}pd^{D}e\int_0^{\infty}dc~exp(i(p_m\Delta x^m-e^mp_m\Delta T +(e^me_m-i0)c^{2\over {D-2}})), \qquad \end{equation} where $c$ is an auxiliary variable. Integrating over $p^m$ and $e^m$ one gets \begin{equation} \label{436} Z_D=\int_{0}^{\infty}dc{1\over{c^{D/2}}} exp(i{{\Delta x^m\Delta x_m}\over{2c}} -c0), \qquad D=3,4,6, \end{equation} or $$ Z_D={1\over {(\Delta x^m\Delta x_m-i0)^{{D-2}\over 2}}}, $$ which coincides with the coordinate propagator for the massless bosonic particle in the standard formulation \cite{govaerts}. On the other hand integrating \p{428} only over $e^m$ we get the massless bosonic particle causal propagator in the form $$ Z_D=\int d^Dp{1\over{p^mp_m+i0}}exp(ip_m\Delta x^m). $$ \subsection{Comment on the $D=10$ case} Above we have restricted our consideration to the space--time dimensions 3, 4 and 6. The case of a bosonic twistor--like particle in $D=10$ is much more sophisticated. The Cartan--Penrose representation of a $D=10$ light--like momentum vector is constructed out of a Majorana--Weyl spinor $\lambda^\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta$ which has 16 independent components \begin{equation}\label{pc} P^m=\l\Gamma^m\l. \end{equation} Transformations of $\l^\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta$ which leave \p{pc} invariant take values on an $S^7$-- sphere (see \cite{es,nispach,bfortschr} and references therein). In contrast to the $D=4$ and $D=6$ case, where such transformations belong to the group $U(1)\sim S^1$ \p{224} and $SU(2)\sim S^3$ \p{233}, respectively, $S^7$ is not a Lie group and its corresponding algebra contains structure functions instead of structure constants. Moreover, among the 10 constraints \p{pc} and 16 constraints $P^{(\l)}_\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta=0$ on the momenta conjugate to $x^m$ and $\l^\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta$ $18=10+16-1-7$ (where 7 comes from $S^7$ and 1 corresponds to local $\tau$--reparametrization) are of the second class. They do not form a representation of the Lorentz group and cause the problem for covariant Hamiltonian analysis. One can overcome these problems in the framework of the Lorentz--harmonic formalism (see \cite{bzstr,bpstv} and references therein), where to construct a light--like vector one introduces eight Majorana--Weyl spinors instead of one $\l^\alpha}\def\b{\beta}\def\l{\lambda}\def\g{\gamma}\def\z{\zeta$. Such a spinor matrix takes values in a spinor representation of the double covering group $Spin(1,9)$ of $SO(1,9)$ and satisfies second--class harmonic conditions. The algebra of the constraints in this ``multi--twistor" case is easier to analyze than that with only one commuting spinor involved. The path--integral BRST quantization of the $D=10$ twistor--like particle is in progress. \section {Conclusion} In the present paper the BRST--BFV quantization of the dynamics of massless bosonic particle in $D=3,4,6$ was performed in the twistor--like formulation. To this end the initially mixed system of the first-- and second--class constraints was converted into the system of first--class constraints by extending the initial phase space of the model with auxiliary variables in a Lorentz--covariant way. The conversion procedure (rather than having been a formal trick) was shown to have a meaning of a symmetry transformation which relates different twistor--like formulations of the bosonic particle, corresponding to different gauge choices in the extended phase space. We quantized the model by use of the extended BRST--BFV scheme for the path--integral quantization. As a result we have presented one of the numerous proofs of the equivalence between the twistor--like and conventional formulation of the bosonic particle mechanics. This example demonstrates peculiar features of treating the twistor--like variables within the course of the covariant Hamiltonian analysis and the BRST quantization, which one should take into account when studying more complicated twistor--like systems, such as superparticles and superstrings. \bigskip \noindent {\bf Acknowledgements.} \noindent The authors are grateful to P. Pasti, M. Tonin, D. V. Volkov and V. G. Zima for useful discussion. I. Bandos and D. Sorokin acknowledge partial support from the INTAS and Dutch Government Grant N 94--2317 and the INTAS Grant N 93--493.
proofpile-arXiv_065-427
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Guidelines} It is well known that the effects on the physics of a field, due to a much heavier field coupled to the former, are not detectable at energies comparable to the lighter mass. More precisely the Appelquist-Carazzone (AC) theorem~\cite{ac} states that for a Green's function with only light external legs, the effects of the heavy loops are either absorbable in a redefinition of the bare couplings or suppressed by powers of $k/M$ where $k$ is the energy scale characteristic of the Green's function (presumably comparable to the light mass), and $M$ is the heavy mass. However the AC theorem does not allow to make any clear prediction when $k$ becomes close to $M$ and, in this region one should expect some non-perturbative effect due to the onset of new physics. \par In the following we shall make use of the Wilson's renormalization group (RG) approach to discuss the physics of the light field from the infrared region up to and beyond the mass of the heavy field. Incidentally, the RG technique has been already employed to proof the AC theorem~\cite{girar}. The RG establishes the flow equations of the various coupling constants of the theory for any change in the observational energy scale; moreover the improved RG equations, originally derived by F.J. Wegner and A. Houghton~\cite{weg}, where the mixing of all couplings (relevant and irrelevant) generated by the blocking procedure is properly taken into account, should allow to handle the non-perturbative features arising when the heavy mass threshold is crossed. \par We shall discuss the simple case of two coupled scalar fields and since we are interested in the modifications of the parameters governing the light field, due to the heavy loops, we shall consider the functional integration of the heavy field only. The action at a given energy scale $k$ is \begin{equation} S_k(\phi,\psi)=\int d^4 x~\left ({1\over 2} \partial_\mu \phi \partial^\mu \phi+ {1\over 2} W(\phi,\psi) \partial_\mu \psi \partial^\mu \psi+ U(\phi,\psi) \right ), \label{eq:acteff} \end{equation} with polynomial $W$ and $U$ \begin{equation} U(\phi,\psi)=\sum_{m,n}{{G_{2m,2n} \psi^{2m}\phi^{2n}}\over {(2n)!(2m)!}}; ~~~~~~~~~~~~~~~~~~~~ W(\phi,\psi)=\sum_{m,n}{{H_{2m,2n}\psi^{2m}\phi^{2n}} \over {(2n)!(2m)!}}. \label{eq:svil} \end{equation} Since we want to focus on the light field, which we choose to be $\psi$, we have simply set to 1 the wave function renormalization of $\phi$. In the following we shall analyse the symmetric phase of the theory with vanishing vacuum energy $G_{0,0}=0$. \par We do not discuss here the procedure employed~\cite{jan} to deduce the RG coupled equations for the couplings in Eq.~\ref{eq:svil}, because it is thoroughly explained in the quoted reference. Since it is impossible to handle an infinite set of equations and a truncation in the sums in Eq.~\ref{eq:svil} is required, we keep in the action only terms that do not exceed the sixth power in the fields and their derivatives. Moreover we choose the initial condition for the RG equations at a fixed ultraviolet scale $\Lambda$ where we set $H_{0,0}=1$, $G_{0,4}= G_{2,2}=G_{4,0}=0.1$ and $G_{0,6}=G_{2,4}=G_{4,2}=G_{6,0}= H_{2,0}=H_{0,2}=0$, and the flow of the various couplings is determined as a function of $t=ln \left (k/ \Lambda\right )$, for negative $t$. \begin{figure} \psfig{figure=fig1.ps,height=4.3cm,width=12.cm,angle=90} \caption{ (a): $G_{0,2}(t)/\Lambda^2$ (curve (1)) and $10^6\cdot G_{2,0}(t)/\Lambda^2$ (curve (2)) vs $t=log\left ({{k}/{\Lambda}}\right )$. \break (b): $G_{2,2}(t)$ (1), $G_{0,4}(t)$ (2), $G_{4,0}(t)$ (3) vs $t$. \label{fig:funo}} \end{figure} \par In Fig.~\ref{fig:funo}(a) $G_{0,2}(t)/\Lambda^2$ (curve (1) ) and $10^6 \cdot G_{2,0}(t)/\Lambda^2$ (curve (2)) are plotted. Clearly the heavy and the light masses become stable going toward the IR region and their value at $\Lambda$ has been chosen in such a way that the stable IR values are, $M\equiv\sqrt {G_{0,2}(t=-18)}\sim 10^{-4}\cdot \Lambda$ and $m\equiv\sqrt{G_{2,0}(t=-18)}\sim 2\cdot 10^{-7}\cdot\Lambda$. So, in principle, there are three scales: $\Lambda$, ($t=0$), the heavy mass $M$, ($t\sim -9.2$), the light mass $m$, ($t\sim -16.1$). In Fig.~\ref{fig:funo}(b) the three renormalizable dimensionless couplings are shown; the neat change around $t=-9.2$, that is $k \sim M$, is evident and the curves become flat below this value. The other four non-renormalizable couplings included in $U$ are plotted in Fig.~\ref{fig:fdue}(a), in units of $\Lambda$. Again everything is flat below $M$, and the values of the couplings in the IR region coincide with their perturbative Feynman-diagram estimate at the one loop level; it is easy to realize that they are proportional to $1/M^2$, which, in units of $\Lambda$, is a big number. Thus the large values in Fig.~\ref{fig:fdue}(a) are just due to the scale employed and, since these four couplings for any practical purpose, must be compared to the energy scale at which they are calculated, it is physically significant to plot them in units of the running cutoff $k$: the corresponding curves are displayed in Fig.~\ref{fig:fdue}(b); in this case the couplings are strongly suppressed below $M$. \begin{figure} \psfig{figure=fig2.ps,height=4.3cm,width=12.cm,angle=90} \caption{ (a): $G_{6,0}(t)\cdot \Lambda^2$ (1), $G_{0,6}(t)\cdot \Lambda^2$ (2), $G_{4,2}(t)\cdot \Lambda^2$ (3) and $G_{2,4}(t)\cdot \Lambda^2$ (4) vs $t$.\break (b): $G_{6,0}(t)\cdot k^2$ (1), $G_{0,6}(t)\cdot k^2$ (2), $G_{4,2}(t)\cdot k^2$ (3) and $G_{2,4}(t)\cdot k^2$ (4) vs $t$. \label{fig:fdue}} \end{figure} \par It must be remarked that there is no change in the couplings when the light mass threshold is crossed. This is a consequence of having integrated the heavy field only: in this case one could check directly from the equations ruling the coupling constants flow, that a shift in the initial value $G_{2,0}(t=0)$ has the only effect (as long as one remains in the symmetric phase) of modifying $G_{2,0}(t)$, leaving the other curves unchanged. Therefore the results obtained are independent of $m$ and do not change even if $m$ becomes much larger than $M$. An example of the heavy mass dependence is shown in Fig.~\ref{fig:ftre}(a), where $G_{6,0}(t)$ is plotted, in units of the running cutoff $k$, for three different values of $G_{0,2}(t=0)$, which correspond respectively to $M/\Lambda\sim 2\cdot 10^{-6}$, (1), $M/\Lambda\sim 10^{-4}$, (2) and $M/\Lambda\sim 0.33$, (3). Note, in each curve, the change of slope when the $M$ scale is crossed. $H_{0,0}=1,~~H_{0,2}=0$ is a constant solution of the corresponding equations for these two couplings; on the other hand $H_{2,0}$ is not constant and it is plotted in units of the running cutoff $k$ in Fig.~\ref{fig:ftre}(b), for the three values of $M$ quoted above. \begin{figure} \psfig{figure=fig3.ps,height=4.3cm,width=12.cm,angle=90} \caption{ (a): $G_{6,0}(t)\cdot k^2$ vs $t$ for $M/\Lambda \sim 2\cdot 10^{-6}$ (1), $\sim 10^{-4}$ (2), $\sim 0.33$ (3).\break (b): $H_{2,0}(t)\cdot k^2$ for the three values of $M/\Lambda$ quoted in (a). \label{fig:ftre}} \end{figure} \par In conclusion, according to the AC theorem all couplings are constant at low energies and a change in the UV physics can only shift their values in the IR region. Remarkably, for increasing $t$, no trace of UV physics shows up until one reaches $M$, that acts as a UV cut-off for the low energy physics. Moreover, below $M$, no non-perturbative effect appears due to the non-renormalizable couplings that vanish fastly in units of $k$. Their behavior above $M$ is somehow constrained by the renormalizability condition fixed at $t=0$, as clearly shown in Fig.~\ref{fig:ftre}(a) (3). Finally, the peak of $H_{2,0}$ at $k\sim M$ in Fig.~\ref{fig:ftre}(b), whose width and height are practically unchanged in the three examples, is a signal of non-locality of the theory limited to the region around $M$. \section*{Acknowledgments} A.B. gratefully acknowledges Fondazione A. Della Riccia and INFN for financial support. \section*{References}
proofpile-arXiv_065-428
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} \label{sec:intro} \subsection{The Lyman--alpha forest} \label{sec:laf} Spectroscopic observations towards quasars show a large number of intervening absorption systems. This `forest' of lines is numerically dominated by systems showing only the Lyman--alpha transition --- these absorbers are called Lyman alpha clouds. Earlier work suggests that the clouds are large, highly ionized structures, either pressure confined (eg. Ostriker \& Ikeuchi 1983) or within cold dark matter structures (eg. Miralda--Escud\'e \& Rees 1993; Cen, Miralda--Escud\'{e} \& Ostriker 1994; Petitjean \& M\"{u}cket 1995; Zhang, Anninos \& Norman 1995). However, alternative models do exist: cold, pressure confined clouds (eg. Barcons \& Fabian 1987, but see Rauch et~al. 1993); various shock mechanisms (Vishniac \& Bust 1987, Hogan 1987). Low and medium resolution spectroscopic studies of the forest generally measure the redshift and equivalent width of each cloud. At higher resolutions it is possible to measure the redshift ($z$), H~I column density ($N$, atoms per cm$^2$) and Doppler parameter ($b$, \hbox{\rm km\thinspace s$^{-1}$}). These are obtained by fitting a Voigt profile to the data (Rybicki \& Lightman 1979). Using a list of $N$, $z$ and $b$ measurements, and their associated error estimates, the number density of the population can be studied. Most work has assumed that the density is a separable function of $z$, $N$ and $b$ (Rauch et~al. 1993). There is a local decrease in cloud numbers near the background quasar which is normally attributed to the additional ionizing flux in that region. While this may not be the only reason for the depletion (the environment near quasars may be different in other respects; the cloud redshifts may reflect systematic motions) it is expected for the standard physical models wherever the ionising flux from the quasar exceeds, or is comparable to, the background. Since the generally accepted cloud models are both optically thin to ionizing radiation and highly ionized, it is possible to correct column densities from the observed values to those that would be seen if the quasar were more remote. The simplest correction assumes that the shape of the two incident spectra --- quasar and background --- are similar. In this case the column density of the neutral fraction is inversely proportional to the incident ionizing flux. If the flux from the quasar is known, and the depletion of clouds is measured from observations, the background flux can be determined. By observing absorption towards quasars at different redshifts the evolution of the flux can be measured. Bechtold (1993) summarises earlier measurements of the ionising flux, both locally and at higher redshifts. Recently Loeb \& Eisenstein (1995) have suggested that enhanced clustering near quasars causes this approach to overestimate the background flux. If this is the case then an analysis which can also study the evolution of the effect gives important information. In particular, a decrease in the inferred flux might be expected after the redshift where the quasar population appears to decrease. However, if the postulated clustering enhancement is related to the turn--on of quasars at high redshift, it may conspire to mask any change in the ionizing background. Section \ref{sec:model} describes the model of the population density in more detail, including the corrections to flux and redshift that are necessary for a reliable result. The data used are described in section \ref{sec:data}. In section \ref{sec:errors} the quality of the fit is assessed and the procedure used to calculate errors in the derived parameters is explained. Results are given in section \ref{sec:results} and their implications discussed in section \ref{sec:discuss}. Section \ref{sec:conc} concludes the paper. \section{The Model} \label{sec:model} \subsection{Population Density} \label{sec:popden} The Doppler parameter distribution is not included in the model since it is not needed to determine the ionizing background from the proximity effect. The model here assumes that $N$ and $z$ are uncorrelated. While this is unlikely (Carswell 1995), it should be a good approximation over the restricted range of column densities considered here. The model of the population without Doppler parameters or the correction for the proximity effect is \begin{equation} dn(N^\prime,z) = \, A^\prime (1+z)^{\gamma^\prime}\,(N^\prime)^{-\beta} \frac{c(1+z)}{H_0(1+2q_0z)^\frac{1}{2}} \ dN^\prime\,dz \end{equation} where $H_0$ is the Hubble parameter, $q_0$ is the cosmological deceleration parameter and $c$ is the speed of light. Correcting for the ionizing flux and changing from `original' ($N^\prime$) to `observed' ($N$) column densities, gives \begin{equation} dn(N,z) = \, A (1+z)^{\gamma^\prime} \left(\frac{N}{\Delta_F}\right)^{-\beta} \frac{c(1+z)}{H_0(1+2q_0z)^\frac{1}{2}} \ \frac{dN}{\Delta_F}\,dz \end{equation} where \begin{equation}N = N^\prime \Delta_F \ , \end{equation} \begin{equation}\Delta_F = \frac{ f_\nu^B }{ f_\nu^B + f_\nu^Q } \ , \end{equation} and $f_\nu^B$ is the background flux, $f_\nu^Q$ is the flux from the quasar ($4\pi J_\nu(z)$). The background flux $J_\nu^B$ may vary with redshift. Here it is parameterised as a constant, a power law, or two power laws with a break which is fixed at $z_B=3.25$ (the mid--point of the available data range). An attempt was made to fit models with $z_B$ as a free parameter, but the models were too poorly constrained by the data to be useful. \begin{eqnarray} J_\nu(z)=10^{J_{3.25}} & \hbox{model {\bf B}} \\ J_\nu(z)=10^{J_{3.25}}\left(\frac{1+z}{1+3.25}\right)^{\alpha_1} & \mbox{{\bf C}} \\ J_\nu(z)=10^{J_{z_B}}\times\left\{\begin{array}{ll}\left(\frac{1+z}{1+z_B}\right)^{\alpha_1}&\mbox{$z<z_B$}\\ \left(\frac{1+z}{1+z_B}\right)^{\alpha_2}&\mbox{$z>z_B$}\end{array}\right. & \mbox{{\bf D \& E}} \end{eqnarray} A large amount of information (figure \ref{fig:nz}) is used to constrain the model parameters. The high--resolution line lists give the column density and redshift, with associated errors, for each line. To calculate the background ionising flux the quasar luminosity and redshift must be known (table \ref{tab:objects}). Finally, each set of lines must have observational completeness limits (table \ref{tab:compl}). \subsection{Malmquist Bias and Line Blending} \label{sec:malm} Malmquist bias is a common problem when fitting models to a population which increases rapidly at some point (often near an observational limit). Errors during the observations scatter lines away from the more populated regions of parameter space and into less populated areas. Line blending occurs when, especially at high redshifts, nearby, overlapping lines cannot be individually resolved. This is a consequence of the natural line width of the clouds and cannot be corrected with improved spectrographic resolution. The end result is that weaker lines are not detected in the resultant `blend'. Both these effects mean that the observed population is not identical to the `underlying' or `real' distribution. \subsubsection{The Idea of Data Quality} To calculate a correction for Malmquist bias we need to understand the significance of the error estimate since any correction involves understanding what would happen if the `same' error occurs for different column density clouds. The same physical cloud cannot be observed with completely different parameters, but the same combination of all the complex factors which influence the errors might affect a line with different parameters in a predictable way. If this idea of the `quality' of an observation could be quantified it would be possible to correct for Malmquist bias: rather than the `underlying' population, one reflecting the quality of the observation (ie. including the bias due to observational errors) could be fitted to the data. For example, if the `quality' of an observation was such that, whatever the actual column density measured, the error in column density was the same, then it would be trivial to convolve the `underlying' model with a Gaussian of the correct width to arrive at an `observed' model. Fitting the latter to the data would give parameters unaffected by Malmquist bias. Another example is the case of galaxy magnitudes. The error in a measured magnitude is a fairly simple function of source brightness, exposure time, etc., and so it is possible to correct a flux--limited galaxy survey for Malmquist bias. \subsubsection{Using Errors as a Measure of Quality} It may be possible to describe the quality of a spectrum by the signal to noise level in each bin. From this one could, for a given line, calculate the expected error in the equivalent width. The error in the equivalent width might translate, depending on whether the absorption line was in the linear or logarithmic portion of the `curve of growth', to a normal error in either $N$ or $\log(N)$. But in this idealised case it has been assumed that the spectrum has not been re--binned, leaving the errors uncorrelated; that the effect of overlapping, blended lines is unimportant; that there is a sudden transition from a linear to logarithmic curve of growth; that the resulting error is well described by a normal distribution. None of this is likely to be correct and, in any case, the resulting analysis, with different `observed' populations for every line, would be too unwieldy to implement, given current computing facilities. A more pragmatic approach might be possible. A plot of the distribution of errors with column density (figure~\ref{fig:ndn}) suggests that the errors in $\log(N)$ are of a similar magnitude for a wide range of lines (although there is a significant correlation between the two parameters). Could the error in $\log(N)$ be a sufficiently good indicator of the `quality' of an observation? \begin{figure} \epsfxsize=8.5cm \epsfbox{nn.ps} \epsfverbosetrue \caption{The distribution of errors in column density.} \label{fig:ndn} \end{figure} If the number density of the underlying population is $n^\prime(N)\ \hbox{d}\log(N)$ then the observed population density for a line with error in $\log(N)$ of $\sigma_N$ is: \begin{equation} n(N)\ \hbox{d}\log(N) \propto\int_{-\infty}^{\infty}n^\prime\left(N10^x\right)\,\exp\left(\frac{-x^2}{2\sigma_N^2}\right)\,\hbox{d}x\ . \end{equation} For a power law distribution this can be calculated analytically and gives an increased probability of seeing lines with larger errors, as expected. For an underlying population density $N^{-\beta}\ \hbox{d}N$ the increase is $\exp\left((1-\beta)^2(\sigma_N\log 10)^2/2\right)$. This gives a lower statistical weight to lines with larger errors when fitting. For this case --- a power law and log--normal errors --- the weighting is not a function of $N$ directly, which might imply that any correction would be uniform, with little bias expected for estimated parameters. In practice this correction does not work. This is probably because the exponential dependence of the correction on $\sigma_N$ makes it extremely sensitive to the assumptions made in the derivation above. These assumptions are not correct. For example, it seems that the correlation between \hbox{$\log( N )$}\ and the associated error is important. \subsubsection{An Estimation of the Malmquist Bias} It is possible to do a simple numerical simulation to gauge the magnitude of the effect of Malmquist bias. A population of ten million column densities were selected at random from a power law distribution ($\beta=1.5, \log(N_{\hbox{min}})=10.9, \log(N_{\hbox{max}})=22.5$) as an `unbiased' sample. Each line was given an `observed' column density by adding a random error distributed with a normal or log--normal (for lines where $13.8 < \log(N) < 17.8$) distribution, with a mean of zero and a standard deviation in $\log(N)$ of $0.5$. This procedure is a simple approximation to the type of errors coming from the curve of growth analysis discussed above, assuming that errors are approximately constant in $\log(N)$ (figure~\ref{fig:ndn}). The size of the error is larger than typical, so any inferred change in $\beta$ should be an upper limit. Since a power--law distribution diverges as $N\rightarrow0$ a normal distribution of errors in $N$ would give an infinite number of observed lines at every column density. This is clearly unphysical (presumably the errors are not as extended as in a normal distribution and the population has some low column density limit). Because of this the `normal' errors above were actually constrained to lie within 3 standard deviations of zero. \begin{figure} \epsfxsize=8.5cm \epsfbox{malm.ps} \epsfverbosetrue \caption{A model including Malmquist bias. The bold, solid line is the original sample, the bold, dashed line is the distribution after processing as described in the text. Each curve is shown twice, but the upper right plot has both axes scaled by a factor of 3 and only shows data for $12.5<\log(N)<16$. Reference lines showing the evolution expected for $\beta=1.5$ and $1.45$ (dashed) are also shown (centre).} \label{fig:malm} \end{figure} The results (figure~\ref{fig:malm}) show that Malmquist bias has only a small effect, at least for the model used here. The main solid line is the original sample, the dashed line is the observed population. Note that the results in this paper come from fitting to a sample of lines with $12.5<\log(N)<16$ (section~\ref{sec:data}) --- corresponding to the data shown expanded to the upper right of the figure. Lines with smaller column densities are not shown since that fraction of the population is affected by the lower density cut--off in the synthetic data. A comparison with the two reference lines, showing the slopes for a population with $\beta=1.5$ or $1.45$ (dashed), indicates that the expected change in $\beta$ is $\sim0.05$. The population of lines within the logarithmic region of the curve of growth appears to be slightly enhanced, but otherwise the two curves are remarkably similar. The variations at large column densities are due to the small number of strong lines in the sample. \subsubsection{Other Approaches} What other approaches can be used to measure or correct the effects of Malmquist bias and line blending? Press \& Rybicki (1993) used a completely different analysis of the Lyman--$\alpha$ forest. Generating and reducing synthetic data, with a known background and cloud population, would allow us to assess the effect of blending. Changing (sub--setting) the sample of lines that is analysed will alter the relative (and, possibly, absolute) importance of the two effects. The procedure used by Press \& Rybicki (1993) is not affected by Malmquist bias or line blending, but it is difficult to adapt to measure the ionizing background. Profile fitting to high--resolution data is a slow process, involving significant manual intervention (we have tried to automate profile--fitting with little success). An accurate measurement of the systematic error in the ionizing background would need an order of magnitude more data than is used here to get sufficiently low error limits. Even if this is possible --- the analysis would need a prohibitive amount of CPU time --- it would be sufficient work for a separate, major paper (we would be glad to supply our software to anyone willing to try this). Taking a sub--set of the data is not helpful unless it is less likely to be affected by the biases described above. One approach might be to reject points with large errors, or large relative errors, in column density since these are more affected by Malmquist bias. However, this would make the observations incomplete in a very poorly understood way. For example, relative errors are correlated with column density (as noted above) and so rejecting lines with larger relative errors would preferentially reject higher column density lines. There is no sense in trying to measure one bias if doing so introduces others. Unlike rejecting lines throughout the sample, changing the completeness limit does not alter the coverage of the observations (or rather, it does so in a way that is understood and corrected for within the analysis). Raising the completeness limits should make line blending less important since weaker lines, which are most likely to be blended, are excluded from the fit. Whether it affects the Malmquist bias depends on the distribution of errors. For blended lines, which tend to be weak, raising the completeness limit should increase the absolute value of $\beta$ since the more populous region of the (hypothetical) power--law population of column densities will no longer be artificially depleted. The effect on $\gamma$ is more difficult to assess since it is uncertain whether the completeness limits are correct at each redshift. If the limits increase too rapidly with redshift, for example, then raising them further will reduce blending most at lower redshifts, lowering $\gamma$. But if they are increasing too slowly then the converse will be true. \subsubsection{Conclusions} Until either profile--fitting is automated, or the method of Press \& Rybicki (1993) can be modified to include the proximity effect, these two sources of uncertainty --- Malmquist bias and line blending --- will continue be a problem for any analysis of the Lyman--$\alpha$ forest. However, from the arguments above, it is likely that the effect of Malmquist bias is small and that, by increasing the completeness limit, we can assess the magnitude of the effect of line blending. \subsection{Flux Calculations} \label{sec:fluxcal} \subsubsection{Galactic Extinction} Extinction within our Galaxy reduces the apparent luminosity of the quasars and so lowers the estimate of the background. Since the absorption varies with frequency this also alters the observed spectral slope. Observed fluxes were corrected using extinction estimates derived from the H~I measurements of Heiles \& Cleary (1979) for Q2204--573 and Stark et~al. (1992) for all the other objects. H I column densities were converted to $E(B-V)$ using the relationships: \begin{eqnarray} E(B-V)&=&\frac{N_{\hbox{\footnotesize H~I}}}{5.27\,10^{21}}\quad\hbox{if\ } \frac{N_{\hbox{\footnotesize H~I}}}{5.27\,10^{21}}<0.1\\ E(B-V)&=&\frac{N_{\hbox{\footnotesize H~I}}}{4.37\,10^{21}}\quad\hbox{otherwise} \end{eqnarray} where the first value comes from Diplas \& Savage (1994) and the second value, which compensates for the presence of H$_2$, is the first scaled by the ratio of the conversions given in Bohlin, Savage \& Drake (1978). A ratio $R=A(V)/E(B-V)$ of 3.0 (Lockman \& Savage 1995) was used and variations of extinction with frequency, $A(\lambda)/A(V)$ were taken from Cardelli, Clayton \& Mathis (1989). The correction to the observed index, $\alpha_o$, of the power--law continuum, \begin{equation} f_\nu\propto\nu^{-\alpha}\ , \end{equation} was calculated using \begin{equation} \alpha_o=\alpha+\frac{A(V)}{2.5}\frac{\partial}{\partial\ln\nu}\frac{A(\nu)}{A(V)} \end{equation} which, using the notation of Cardelli, Clayton \& Mathis (1989), becomes \begin{equation} \alpha_o=\alpha+\frac{A(V)}{2.5\,10^6c}\,\nu\ln(10)\,\frac{\partial}{\partial y}\left(a(x)+\frac{b(x)}{R}\right)\ . \end{equation} \subsubsection{Extinction in Damped Systems} \label{sec:dampcor} Two quasars are known to have damped absorption systems along the line of sight (Wolfe et~al. 1995). The extinction due to these systems is not certain, but model {\bf E} includes the corrections listed in table~\ref{tab:damp}. These have been calculated using the SMC extinction curve in Pei (1992), with a correction for the evolution of heavy element abundances taken from Pei \& Fall (1995). The SMC extinction curve is most suitable for these systems since they do not appear to have structure at 2220~\AA\ (Boiss\'{e} \& Bergeron 1988), unlike LMC and Galactic curves. \begin{table} \begin{tabular}{lllll} Object&$\log(N_{\hbox{\footnotesize H I}})$&$z_{\hbox{\footnotesize abs}}$&$A(V)$&$\Delta_\alpha$\\ Q0000--263&21.3&3.39&0.10 &0.078\\ Q2206--199&20.7&1.92&0.14 &0.10\\ $ $&20.4&2.08&0.019 &0.049\\ \end{tabular} \caption{ The damped absorption systems and associated corrections (at 1450~\AA\ in the quasar's rest--frame) for model {\bf E}.} \label{tab:damp} \end{table} \subsubsection{Absorption by Clouds near the Quasar} \label{sec:internal} The amount of ionizing flux from the background quasar incident on a cloud is attenuated by all the other clouds towards the source. If one of the intervening clouds has a large column density this can significantly reduce the extent of the effect of the quasar. To correct for this the fraction of ionizing photons from the quasar not attenuated by the intervening H~I and He~II absorption is estimated before fitting the model. A power--law spectrum is assumed and the attenuation is calculated for each cloud using the cross--sections given in Osterbrock (1989). The ratio $n(\hbox{He~II})/n(\hbox{H~I})$ within the clouds will depend on several unknown factors (the true energy distribution of the ionizing flux, cloud density, etc.), but was assumed to be 10 (Sargent et~al. 1980, Miralda--Escud\'e 1993). The attenuation is calculated using all the observed intervening clouds. This includes clouds which are not included in the main fit because they lie outside the column density limits, or are too close to the quasar ($\Delta z \leq 0.003$). For most clouds ($\log(N)\sim13.5$) near enough to the quasar to influence the calculation of the background this correction is unimportant (less than 1\%). However, large ($\log(N)\sim18$ or larger) clouds attenuate the flux to near zero. This explains why clouds with $\Delta_f\sim1$ are apparent close to the QSO in figure~\ref{fig:prox}. This relatively sudden change in optical depth at $\log(N)\sim18$ is convenient since it makes the correction insensitive to any uncertainties in the calculation (eg. $n(\hbox{He~II})/n(\hbox{H~I})$, the shape of the incident spectrum, absorption by heavier elements) --- for most column densities any reasonable model is either insignificant ($\log(N)<17$) or blocks practically all ionizing radiation ($\log(N)>19$). In fact, the simple correction described above is in reasonable agreement with CLOUDY models, for even the most critical column densities. A model cloud with a column density of $\log(N)=13.5$ and constant density was irradiated by an ionizing spectrum based on that of Haardt \& Madau (1996). Between the cloud and quasar the model included an additional absorber (constant density, $\log(N)=18$) which modified the quasar's spectrum. The effect of the absorber (for a range of heavy element abundances from pure H to primordial to 0.1 solar) on the ionized fraction of H~I was consistent with an inferred decrease in the quasar flux of about 80\%. In comparison, the correction above, using a power--law spectrum with $\alpha=1$, gave a reduction of 60\% in the quasar flux. These two values are in good agreement, considering the exponential dependence on column densities and the uncertainty in spectral shape. At higher and lower absorber column densities the agreement was even better, as expected. \subsection{Redshift Corrections} \label{sec:redcor} Gaskell (1982) first pointed out a discrepancy between the redshifts measured from Lyman $\alpha$ and C~IV emission, and those from lower ionization lines (eg. Mg~II, the Balmer series). Lower ionization lines have a larger redshift. If the systemic redshift of the quasar is assumed to be that of the extended emission (Heckman et~al. 1991), molecular emission (Barvainis et~al. 1994), or forbidden line emission (Carswell et~al. 1991), then the low ionization lines give a better measure of the rest--frame redshift. Using high ionization lines gives a reduced redshift for the quasar, implies a higher incident flux on the clouds from the quasar, and, for the same local depletion of lines, a higher estimate of the background. Espey (1993) re--analysed the data in Lu, Wolfe \& Turnshek (1991), correcting systematic errors in the quasar redshifts. The analysis also considered corrections for optically thick and thin universes and the differences between the background and quasar spectra, but the dominant effect in reducing the estimate from 174 to 50~J$_{23}$\ was the change in the quasar redshifts. To derive a more accurate estimate of the systemic velocity of the quasars in our sample we made use of published redshift measurements of low ionization lines, or measured these where spectra were available to us. The lines used depended on the redshift and line strengths in the data, but typically were one or more of Mg~II$\,2798\,$\AA, O~I$\,1304\,$\AA, and C~II$\,1335\,$\AA. When no low ionization line observations were available (Q0420--388, Q1158--187, Q2204--573) we applied a mean correction to the high ionization line redshifts. These corrections are based on measurements of the relative velocity shifts between high and low ionization lines in a large sample of quasars (Espey \& Junkkarinen 1996). They find a correlation between quasar luminosity and mean velocity difference ($\Delta_v$) with an empirical relationship given by: \begin{equation} \Delta_v=\exp(0.66\log L_{1450}-13.72)\ \hbox{\rm km\thinspace s$^{-1}$}\end{equation} where $L_{1450}$ is the rest--frame luminosity (ergs Hz$^{-1}$ s$^{-1}$) of the quasar at 1450~\AA\ for $q_0=0.5$ and H$_0=100\ \hbox{\rm km\thinspace s$^{-1}$}/\hbox{Mpc}$. \section{The Data} \label{sec:data} \begin{table*} \begin{tabular}{lrrrrrrrr} &&&\multicolumn{2}{c}{$L_\nu(1450)$}&\multicolumn{4}{c}{Typical change in $\log(\hbox{J$_{23}$})$}\\ \hfill Object \hfill&\hfill $z$ \hfill&\hfill $\alpha$ \hfill&\hfill $q_0=0$ \hfill&$\hfill q_0=0.5$ \hfill&\hfill $z$ \hfill&\hfill $f_\nu$ \hfill&\hfill $\alpha$ \hfill&\hfill Total \hfill\\ Q0000--263 & 4.124 & 1.02 & 13.5\ten{31} & 2.8\ten{31} & $-$0.09 & $+$0.02 & $+$0.01 & $-$0.05\\ Q0014+813 & 3.398 & 0.55 & 34.0\ten{31} & 8.6\ten{31} & $-$0.19 & $+$0.33 & $+$0.21 & $+$0.36\\ Q0207--398 & 2.821 & 0.41 & 5.6\ten{31} & 1.7\ten{31} & $-$0.16 & $+$0.03 & $+$0.02 & $-$0.11\\ Q0420--388 & 3.124 & 0.38 & 10.9\ten{31} & 3.0\ten{31} & $-$0.16 & $+$0.04 & $+$0.02 & $-$0.10\\ Q1033--033 & 4.509 & 0.46 & 5.5\ten{31} & 1.0\ten{31} & $-$0.05 & $+$0.12 & $+$0.00 & $+$0.06\\ Q1100--264 & 2.152 & 0.34 & 13.8\ten{31} & 5.3\ten{31} & $-$0.42 & $+$0.19 & $+$0.11 & $-$0.13\\ Q1158--187 & 2.454 & 0.50 & 42.2\ten{31} & 14.4\ten{31} & $-$0.46 & $+$0.09 & $+$0.06 & $-$0.31\\ Q1448--232 & 2.223 & 0.61 & 9.6\ten{31} & 3.5\ten{31} & $-$0.34 & $+$0.28 & $+$0.17 & $+$0.11\\ Q2000--330 & 3.783 & 0.85 & 12.7\ten{31} & 2.9\ten{31} & $-$0.10 & $+$0.16 & $+$0.10 & $+$0.15\\ Q2204--573 & 2.731 & 0.50 & 42.8\ten{31} & 13.3\ten{31} & $-$0.35 & $+$0.06 & $+$0.03 & $-$0.25\\ Q2206--199 & 2.574 & 0.50 & 19.4\ten{31} & 6.3\ten{31} & $-$0.31 & $+$0.05 & $+$0.03 & $-$0.22\\ Mean & 3.081 & 0.56 & 19.1\ten{31} & 5.7\ten{31} & $-$0.24 & $+$0.12 & $+$0.07 & $-$0.04\\ \end{tabular} \caption{ The systemic redshifts, power law continuum exponents. and rest frame luminosities (ergs Hz$^{-1}$ s$^{-1}$\ at 1450~\AA) for the quasars used. $H_0=100$ \hbox{\rm km\thinspace s$^{-1}$}/Mpc and luminosity scales as $H_0^{-2}$. The final four columns are an estimate of the relative effect of the various corrections in the paper (systemic redshift, correction for reddening to flux and spectral slope).} \label{tab:objects} \end{table*} Objects, redshifts and fluxes are listed in table \ref{tab:objects}. A total of 1675 lines from 11 quasar spectra were taken from the literature. Of these, 844 lie within the range of redshifts and column densities listed in table \ref{tab:compl}, although the full sample is used to correct for absorption between the quasar and individual clouds (section~\ref{sec:internal}). The lower column density limits are taken from the references; upper column densities are fixed at $\hbox{$\log( N )$}=16$ to avoid the double power law distribution discussed by Petitjean et~al. (1993). Fluxes are calculated using standard formulae, assuming a power law spectrum ($f_\nu \propto \nu^{-\alpha}$), with corrections for reddening. Low ionization line redshifts were used where possible, otherwise high ionization lines were corrected using the relation given in section~\ref{sec:redcor}. Values of $\alpha$ uncorrected for absorption are used where possible, corrected using the relation above. If no $\alpha$ was available, a value of 0.5 was assumed (Francis 1993). References and notes on the calculations for each object follow: \begin{description} \item[Q0000--263] Line list from Cooke (1994). There is some uncertainty in the wavelength calibration for these data, but the error ($\sim30 \hbox{\rm km\thinspace s$^{-1}$}$) is much less than the uncertainty in the quasar redshift ($\sim900 \hbox{\rm km\thinspace s$^{-1}$}$) which is taken into account in the error estimate (section~\ref{sec:erress}). Redshift this paper (section~\ref{sec:redcor}). Flux and $\alpha$ measurements from Sargent, Steidel \& Boksenberg (1989). \item[Q0014+813] Line list from Rauch et~al. (1993). Redshift this paper (section~\ref{sec:redcor}). Flux and $\alpha$ measurements from Sargent, Steidel \& Boksenberg (1989). \item[Q0207--398] Line list from Webb (1987). Redshift (O I line) from Wilkes (1984). Flux and $\alpha$ measurements from Baldwin et~al. (1995). \item[Q0420--388] Line list from Atwood, Baldwin \& Carswell (1985). Redshift, flux and $\alpha$ from Osmer (1979) (flux measured from plot). The redshifts quoted in the literature vary significantly, so a larger error (0.01) was used in section~\ref{sec:erress}. \item[Q1033--033] Line list and flux from Williger et~al. (1994). From their data, $\alpha=0.78$, without a reddening correction. Redshift this paper (section~\ref{sec:redcor}). \item[Q1100--264] Line list from Cooke (1994). Redshift from Espey et~al. (1989) and $\alpha$ from Tytler \& Fan (1992). Flux measured from Osmer \& Smith (1977). \item[Q1158--187] Line list from Webb (1987). Redshift from Kunth, Sargent \& Kowal (1981). Flux from Adam (1985). \item[Q1448--232] Line list from Webb (1987). Redshift from Espey et~al. (1989). Flux and $\alpha$ measured from Wilkes et~al. (1983), although a wide range of values exist in the literature and so a larger error (0.6 magnitudes in the flux) was used in section~\ref{sec:erress}. \item[Q2000--330] Line list from Carswell et~al. (1987). Redshift this paper (section~\ref{sec:redcor}). Flux and $\alpha$ measurements from Sargent, Steidel \& Boksenberg (1989). \item[Q2204--573] Line list from Webb (1987). Redshift from Wilkes et~al. (1983). V magnitude from Adam (1985). \item[Q2206--199] Line list from Rauch et~al. (1993). Redshift this paper (section~\ref{sec:redcor}). V magnitude from Hewitt \& Burbidge (1989). \end{description} \begin{table} \begin{tabular}{cccccc} Object&\multispan2{\hfil$N$\hfil}&\multispan2{\hfil$z$\hfil}&Number\\ name&Low&High&Low&High&of lines\\ Q0000--263& 14.00 & 16.00 & 3.1130 & 3.3104 & 62 \\ & & & 3.4914 & 4.1210 & 101 \\ Q0014+813& 13.30 & 16.00 & 2.7000 & 3.3800 & 191 \\ Q0207--398& 13.75 & 16.00 & 2.0765 & 2.1752 & 11 \\ & & & 2.4055 & 2.4878 & 7 \\ & & & 2.6441 & 2.7346 & 6 \\ & & & 2.6852 & 2.7757 & 9 \\ & & & 2.7346 & 2.8180 & 8 \\ Q0420--388& 13.75 & 16.00 & 2.7200 & 3.0800 & 73 \\ Q1033--033& 14.00 & 16.00 & 3.7000 & 3.7710 & 16 \\ & & & 3.7916 & 3.8944 & 21 \\ & & & 3.9191 & 4.0301 & 24 \\ & & & 4.0548 & 4.1412 & 25 \\ & & & 4.1988 & 4.3139 & 30 \\ & & & 4.3525 & 4.4490 & 23 \\ & & & 4.4517 & 4.4780 & 2 \\ Q1100--264& 12.85 & 16.00 & 1.7886 & 1.8281 & 2 \\ & & & 1.8330 & 1.8733 & 8 \\ & & & 1.8774 & 1.9194 & 13 \\ & & & 1.9235 & 1.9646 & 9 \\ & & & 1.9696 & 2.0123 & 10 \\ & & & 2.0189 & 2.0617 & 6 \\ & & & 2.0683 & 2.1119 & 18 \\ Q1158--187& 13.75 & 16.00 & 2.3397 & 2.4510 & 9 \\ Q1448--232& 13.75 & 16.00 & 2.0847 & 2.1752 & 9 \\ Q2000--330& 13.75 & 16.00 & 3.3000 & 3.4255 & 23 \\ & & & 3.4580 & 3.5390 & 15 \\ & & & 3.5690 & 3.6440 & 18 \\ & & & 3.6810 & 3.7450 & 11 \\ Q2204--573& 13.75 & 16.00 & 2.4467 & 2.5371 & 10 \\ & & & 2.5454 & 2.6276 & 12 \\ & & & 2.6441 & 2.7280 & 8 \\ Q2206--199& 13.30 & 16.00 & 2.0864 & 2.1094 & 2 \\ & & & 2.1226 & 2.1637 & 8 \\ & & & 2.1760 & 2.2188 & 5 \\ & & & 2.2320 & 2.2739 & 7 \\ & & & 2.2887 & 2.3331 & 7 \\ & & & 2.3471 & 2.3940 & 10 \\ & & & 2.4105 & 2.4574 & 4 \\ & & & 2.4754 & 2.5215 & 11 \\ \multicolumn{2}{l}{Total: 11 quasars }& & & & 844 \\ \end{tabular} \caption{Completeness limits.} \label{tab:compl} \end{table} Table~\ref{tab:objects} also gives an estimate of the relative effect of the different corrections made here. Each row gives the typical change in $\log(\hbox{J$_{23}$})$ that would be estimated using that quasar alone, with a typical absorption cloud 2~Mpc from the quasar ($q_0=0.5, H_0=100\,\hbox{\hbox{\rm km\thinspace s$^{-1}$}}$). The correction to obtain the systemic redshift is not necessary for any quasar whose redshift has been determined using low ionization lines. In such cases the value given is the expected change if the redshift measurement had not been available. Using the systematic redshift always reduces the background estimate, while correcting for reddening always acts in the opposite sense. The net result, in the final column of table~\ref{tab:objects}, depends on the relative strength of these two factors. For most objects the redshift correction dominates, lowering $\log(\hbox{J$_{23}$})$ by $\sim 0.15$ (a decrease of 30\%), but for four objects the reddening is more important (Q0014+813, the most reddened, has $B-V = 0.33$; the average for all other objects is $0.09$). \begin{figure*} \hbox{ \epsfxsize=8.5cm \epsfbox{nz0.ps} \hfill \epsfxsize=8.5cm \epsfbox{nz1.ps} } \epsfverbosetrue \caption{ The lines in the full sample (left) used to calculate the attenuation of the quasar flux by intervening clouds and the restricted sample (right) to which the model was fitted.} \label{fig:nz} \end{figure*} Figure \ref{fig:nz} shows the distribution of column density, $N$, and redshift, $z$, for the lines in the sample. The completeness limit was taken from the literature and depends on the quality of the spectra. There is also a clear trend with redshift as the number density increases and weak lines become less and less easy to separate in complex blends, whatever the data quality (see section~\ref{sec:malm} for a more detailed discussion of line blending). \section{Fit Quality and Error Estimates} \label{sec:errors} \subsection{The Quality of the Fit} \label{sec:finalq} \begin{table} \begin{tabular}{c@{\hspace{3em}}cc@{\hspace{3em}}cc} &\multicolumn{2}{c}{Without Inv. Eff.\hfill}&\multicolumn{2}{c}{With Inv. Eff.\hfill}\\ Variable&Statistic&Prob.&Statistic&Prob.\\ $N$ & 1.11 & 0.17 & 1.05 & 0.22 \\ $z$ & 1.12 & 0.16 & 1.05 & 0.22 \\ \end{tabular} \caption{ The K--S statistics measuring the quality of the fit.} \label{tab:ks} \end{table} Figures \ref{fig:cum1} and \ref{fig:cum2} show the cumulative data and model for each variable using two models: one includes the proximity effect (model {\bf B}), one does not (model {\bf A}). The probabilities of the associated K--S statistics are given in table \ref{tab:ks}. For the column density plots the worst discrepancy between model and data occurs at $\log(N)=14.79$. The model with the proximity effect (to the right) has slightly more high column density clouds, as would be expected, although this is difficult to see in the figures (note that the dashed line --- the model --- is the curve that has changed). In the redshift plots the difference between the two models is more apparent because the changes are confined to a few redshifts, near the quasars, rather than, as in the previous figures, spread across a wide range of column densities. The apparent difference between model and data is larger for the model that includes the proximity effect (on the right of figure~\ref{fig:cum2}). However, this is an optical illusion as the eye tends to measure the vertical difference between horizontal, rather than diagonal, lines. In fact the largest discrepancy in the left figure is at $z=3.323$, shifting to $z=3.330$ when the proximity effect is included. It is difficult to assess the importance of individual objects in cumulative plots, but the main difference in the redshift figure occurs near the redshift of Q0014+813. However, since this is also the case without the proximity effect (the left--hand figure) it does not seem to be connected to the unusually large flux correction for this object (section~\ref{sec:data}). In both cases --- with and without the proximity effect --- the model fits the data reasonably well. It is not surprising that including the proximity effect only increases the acceptability of the fit slightly, as the test is dominated by the majority of lines which are not influenced by the quasar. The likelihood ratio test that we use in section \ref{sec:disevid} is a more powerful method for comparing two models, but can only be used if the models are already a reasonable fit (as shown here). \begin{figure*} \hbox{ \epsfxsize=8.5cm \epsfbox{ks0.n.ps} \hfill \epsfxsize=8.5cm \epsfbox{ks1.n.ps} } \epsfverbosetrue \caption{ The cumulative data (solid line) and model (dashed), integrating over $z$, for the lines in the sample, plotted against column density (\hbox{$\log( N )$}). The model on the right includes the proximity effect.} \label{fig:cum1} \end{figure*} \begin{figure*} \hbox{ \epsfxsize=8.5cm \epsfbox{ks0.z.ps} \hfill \epsfxsize=8.5cm \epsfbox{ks1.z.ps} } \epsfverbosetrue \caption{ The cumulative data (solid line) and model (dashed), integrating over $N$, for the lines in the sample, plotted against redshift. The model on the right includes the proximity effect.} \label{fig:cum2} \end{figure*} \subsection{Sources of Error} \label{sec:erress} There are two sources of stochastic uncertainty in the values of estimated parameters: the finite number of observations and the error associated with each observation (column densities, redshifts, quasar fluxes, etc.). The first source of variation --- the limited information available from a finite number of observations --- can be assessed by examining the distribution of the posterior probabilities for each parameter. This is described in the following section. The second source of variation --- the errors associated with each measurement --- can be assessed by repeating the analysis for simulated sets of data. In theory these errors could have been included in the model and their contribution would have been apparent in the posterior distribution. In practice there was insufficient information or computer time to make a detailed model of the error distribution. Instead, ten different sets of line--lists were created. Each was based on the original, with each new value, $X$, calculated from the observed value $x$ and error estimate $\sigma_X$: \begin{equation} X = x + a \sigma_X\ ,\end{equation} where $a$ was selected at random from a (approximate) normal distribution with zero mean and unit variance. The redshift (standard error 0.003) and luminosity (standard error 0.2 magnitudes) of each background quasar were also changed. For Q0420--388 the redshift error was increased to 0.1 and, for Q1448--232, the magnitude error was increased to 0.6 magnitudes. The model was fitted to each data set and the most likely values of the parameters recorded. A Gaussian was fitted to the distribution of values. In some cases (eg.\ figure~\ref{fig:alphas_d}) a Gaussian curve may not be the best way to describe the distribution of measurements. However, since the error in the parameters is dominated by the small number of data points, rather than the observational errors, using a different curve will make little difference to the final results. Since the two sources of stochastic error are not expected to be correlated they can be combined to give the final distribution of the parameters. The Gaussian fitted to the variation from measurement errors is convolved with the posterior distribution of the variable. The final, normalized distribution is then a good approximation to the actual distribution of values expected. This procedure is shown in figures \ref{fig:beta_gamma_d}\ to \ref{fig:alphas_d}. For each parameter in the model the `raw' posterior distribution is plotted (thin line and points). The distribution of values from the synthetic data is shown as a dashed histogram and the fitted Gaussian is a thin line. The final distribution, after convolution, is the heavy line. In general the uncertainties due to a finite data set are the main source of error. \subsection{Error Estimates from Posterior Probabilities\label{sec:postprob}} \newcommand{{\bf y}}{{\bf y}} \newcommand{\tb}{{\bf\theta}} \newcommand{\rb}{{\bf R_\nu}} \newcommand{\bn}{{b_\nu}} If $p({\bf y}|\tb)$ is the likelihood of the observations (${\bf y}$), given the model (with parameters $\tb$), then we need an expression for the posterior probability of a `parameter of interest', $\eta$. This might be one of the model parameters, or some function of the parameters (such as the background flux at a certain redshift): \begin{equation}\eta = g(\tb)\ .\end{equation} For example, the value of J$_{23}$\ at a particular redshift for models {\bf C} to {\bf E} in section~\ref{sec:popden} is a linear function of several parameters (two or more of $J_{3.25}, J_{z_B}, \alpha_1$, and $\alpha_2$). To calculate how likely a particular flux is the probabilities of all the possible combinations of parameter values consistent with that value must be considered: it is necessary to integrate over all possible values of $\beta$ and $\gamma^\prime$, and all values of $J_{z_B}, \alpha_1$, etc. which are consistent with J$_{23}$(z) having that value. In other words, to find the posterior distribution of $\eta$, $\pi(\eta|\tb)$, we must marginalise the remaining model parameters: \begin{equation}\pi(\eta|\tb)=\lim_{\gamma \rightarrow 0} \frac{1}{\gamma} \int_D \pi(\tb|{\bf y})\,d\tb\ ,\end{equation} where D is the region of parameter space for which $\eta \leq g(\tb) \leq \eta + \gamma$ and $\pi(\tb|{\bf y}) \propto \pi(\tb) p({\bf y}|\tb)$, the posterior density of $\tb$ with prior $\pi(\tb)$. A uniform prior is used here for all parameters (equivalent to normal maximum likelihood analysis). Explicitly, power law exponents and the logarithm of the flux have prior distributions which are uniform over $[-\infty,+\infty]$. Doing the multi--dimensional integral described above would require a large (prohibitive) amount of computer time. However, the log--likelihood can be approximated by a second order series expansion in $\tb$. This is equivalent to assuming that the other parameters are distributed as a multivariate normal distribution, and the result can then be calculated analytically. Such a procedure is shown, by Leonard, Hsu \& Tsui (1989), to give the following procedure when $g(\tb)$ is a linear function of $\tb$: \begin{equation}\bar{\pi}(\eta|{\bf y}) \propto \frac{\pi_M(\eta|{\bf y})}{|\rb|^{1/2}(\bn^T\rb^{-1}\bn)^{1/2}}\ ,\end{equation} where \begin{eqnarray} \pi_M(\eta|{\bf y}) & = & \sup_{\tb:g(\tb)=\eta} \pi(\tb|{\bf y})\\&=&\pi(\tb_\eta|{\bf y})\ ,\\ \bn & = & \left.\frac{\partial g(\tb)}{\partial \tb}\right|_{\tb=\tb_\eta}\ ,\\ \rb & = & \left.\frac{\partial^2 \ln \pi(\tb|{\bf y})}{\partial(\tb\tb^T)} \right|_{\tb=\tb_\eta}\ . \end{eqnarray} The likelihood is maximised with the constraint that $g(\tb)$ has a particular value. $\rb$ is the Hessian matrix used in the fitting routine (Press et~al. 1992) and $\bn$ is known (when $\eta$ is the average of the first two of three parameters, for example, $\bn = 0.5,0.5,0$). This quickens the calculation enormously. To estimate the posterior distribution for, say, J$_{23}$, it is only necessary to choose a series of values and, at each point, find the best fit consistent with that value. The Hessian matrix, which is returned by many fitting routines, can then be used --- following the formulae above --- to calculate an approximation to the integral, giving a value proportional to the probability at that point. Once this has been repeated for a range of different values of J$_{23}$\ the resulting probability distribution can be normalised to give an integral of one. Note that this procedure is only suitable when $g(\tb)$ is a linear function of $\tb$ --- Leonard, Hsu \& Tsui (1989) give the expressions needed for more complex parameters. \section{Results} \label{sec:results} A summary of the results for the different models is given in table \ref{tab:fpars}. The models are: \begin{description} \item[{\bf A}] --- No Proximity Effect. The population model described in section \ref{sec:model}, but without the proximity effect. \item[{\bf B}] --- Constant Background. The population model described in section \ref{sec:model} with a constant ionising background. \item[{\bf C}] --- Power Law Background. The population model described in section \ref{sec:model} with an ionising background which varies as a power law with redshift \item[{\bf D}] --- Broken Power Law Background. The population model described in section \ref{sec:model} with an ionising background whose power law exponent changes at $z_B=3.25$. \item[{\bf E}] --- Correction for Extinction in Damped Systems. As {\bf D}, but with a correction for absorption in known damped absorption systems (section \ref{sec:dampcor}). \end{description} In this paper we assume $q_0$ = 0.5 and $H_0$ = 100~\hbox{\rm km\thinspace s$^{-1}$}/Mpc. \subsection{Population Distribution} \label{sec:resmpars} \begin{table*} \begin{tabular}{crlrlrlrlrlrc} Model & \multispan2{\hfill$\beta$\hfill} & \multispan2{\hfill$\gamma$\hfill} & \multispan2{\hfill $J_{z_B}$\hfill} & \multispan2{\hfill $\alpha_1$\hfill} & \multispan2{\hfill $\alpha_2$\hfill} & \multispan1{\hfill $z_B$\hfill} & -2 log--likelihood\\ {\bf A} & 1.66 & $\pm0.03$ & 2.7 & $\pm0.3$ & \multicolumn{7}{c}{No background} & 60086.2 \\ {\bf B} & 1.67 & $\pm0.03$ & 2.9 & $\pm0.3$ & $-21.0$ & $\pm0.2$ & & & & & & 60052.8 \\ {\bf C} & 1.67 & $\pm0.03$ & 3.0 & $\pm0.3$ & $-21.0$ & $\pm0.2$ & $-1$ & $\pm3$ & & & & 60052.6 \\ {\bf D} & 1.67 & $\pm0.04$ & 3.0 & $\pm0.3$ & $-20.9$ & $\pm0.3$ & 0 & $+5,-6$ & $-2$ & $\pm4$ & 3.25 & 60052.4 \\ {\bf E} & 1.67 & $\pm0.03$ & 3.0 & $\pm0.3$ & $-20.9$ & $+0.3,-0.2$ & 0 & $+5,-6$ & $-2$ & $+7,-4$ & 3.25 & 60051.4 \\ \end{tabular} \caption{ The best--fit parameters and expected errors for the models.} \label{tab:fpars} \end{table*} The maximum likelihood `best--fit' values for the parameters are given in table \ref{tab:fpars}. The quoted errors are the differences (a single value if the distribution is symmetric) at which the probability falls by the factor $1/\sqrt{e}$. This is equivalent to a `$1\sigma$ error' for parameters with normal error distributions. The observed evolution of the number of clouds per unit redshift is described in the standard notation found in the literature \begin{equation}dN/dz = A_0 (1+z)^{\gamma}\ .\end{equation} The variable used in the maximum likelihood fits here, $\gamma^\prime$, excludes variations expected from purely cosmological variations and is related to $\gamma$ by: \begin{equation} \gamma = \left\{ \begin{array}{ll}\gamma^\prime + 1 & \mbox{ if $q_0 = 0$} \\ \gamma^\prime + \frac{1}{2} & \mbox{ if $q_0 = 0.5$ \ .}\end{array} \right.\end{equation} Figure \ref{fig:var_d}\ shows the variation in population parameters for model {\bf D} as the completeness limits are increased in steps of $\Delta\hbox{$\log( N )$}=0.1$. The number of clouds decreases from 844 to 425 (when the completeness levels have been increased by $\Delta\hbox{$\log( N )$}=0.5$). \begin{figure*} \hbox{ \epsfxsize=8.5cm \epsfbox{beta.ps} \hfill \epsfxsize=8.5cm \epsfbox{gamma.ps} } \epsfverbosetrue \caption{The expected probability distribution of the model parameters $\beta$ and $\gamma$ (heavy line) for model {\bf D}. The dashed histogram and Gaussian (thin line) show how the measured value varies for different sets of data. The dash-dot line shows the uncertainty in the parameter because the data are limited. These are combined to give the final distribution (bold). See section \ref{sec:erress} for mode details.} \label{fig:beta_gamma_d} \end{figure*} \begin{figure*} \hbox{ \epsfxsize=8.5cm \epsfbox{nu1_b.ps} \hfill \epsfxsize=8.5cm \epsfbox{nu1_d.ps} } \epsfverbosetrue \caption{The expected probability distribution of the log background flux at $z=3.25$ (heavy line) for models {\bf B} (left) and {\bf D}. The uncertainty from the small number of lines near the quasar (line with points) is significantly larger than that from uncertainties in column densities or quasar properties (thin curve). See section \ref{sec:erress} for a full description of the plot.} \label{fig:nu1_d} \end{figure*} \begin{figure*} \hbox{ \epsfxsize=8.5cm \epsfbox{nu2_d.ps} \hfill \epsfxsize=8.5cm \epsfbox{nu3_d.ps} } \epsfverbosetrue \caption{The expected probability distribution of the model parameters $\alpha_1$ and $\alpha_2$ (heavy line) for model {\bf D}. See section \ref{sec:erress} for a full description of the plot.} \label{fig:alphas_d} \end{figure*} \begin{figure*} \hbox{ \epsfxsize=8.5cm \epsfbox{beta_all.ps} \hfill \epsfxsize=8.5cm \epsfbox{gamma_all.ps} } \epsfverbosetrue \caption{The expected probability distribution of the population parameters for model {\bf D}. The top curve is for all data, each lower curve is for data remaining when the column density completeness limits are progressively increased by $\Delta\hbox{$\log( N )$}=0.1$.} \label{fig:var_d} \end{figure*} \subsection{Ionising Background} \label{sec:resbackg} Values of the ionising flux parameters are show in table \ref{tab:fpars}. The expected probability distributions for models {\bf B} and {\bf D} are shown in figures \ref{fig:nu1_d} and \ref{fig:alphas_d}. The background flux relation is described in section \ref{sec:popden}. The variables used to describe the variation of the flux with redshift are strongly correlated. To illustrate the constraints more clearly the marginalised posterior distribution (section \ref{sec:postprob}) of J$_{23}$\ was calculated at a series of redshifts. These are shown (after convolution with the combination of Gaussians appropriate for the uncertainties in the parameters from observational errors) for model {\bf D} in figure \ref{fig:flux_both}. The distribution at each redshift is calculated independently. This gives a conservative representation since the marginalisation procedure assumes that parameters can take all possible values consistent with the background at that redshift (the probability that the flux can be low at a certain redshift, for example, includes the possibility that it is higher at other redshifts). Figure \ref{fig:flux_both} also compares the results from the full data set (solid lines and smaller boxes) with those from the data set with column density completeness limits raised by $\Delta\log(N)=0.5$ (the same data as the final curves in figure \ref{fig:var_d}). Table~\ref{tab:modeld} gives the most likely flux (at probability $p_m$), an estimate of the `1$\sigma$ error' (where the probability drops to $p_m/\sqrt{e}$), the median flux, the upper and lower quartiles, and the 5\% and 95\% limits for model {\bf D} at the redshifts shown in figure~\ref{fig:flux_both}. It is difficult to assess the uncertainty in these values. In general the central measurements are more reliable than the extreme limits. The latter are more uncertain for two reasons. First, the distribution of unlikely models is more likely to be affected by assumptions in section~\ref{sec:postprob} on the normal distribution of secondary parameters. Second, the tails of the probability distribution are very flat, making the flux value sensitive to numerical noise. Extreme limits, therefore, should only be taken as a measure of the relevant flux magnitude. Most likely and median values are given to the nearest integer to help others plot our results --- the actual accuracy is probably lower. \begin{table} \begin{tabular}{crrrrrr} &$z=2$&$z=2.5$&$z=3$&$z=3.5$&$z=4$&$z=4.5$\\ $p_m/\sqrt{e}$ & 30 & 50 & 60 & 60 & 40 & 30 \\ $p_m$ & 137 & 129 & 118 & 103 & 80 & 63 \\ $p_m/\sqrt{e}$ &1000 & 400 & 220 & 180 & 160 & 170 \\ 5\% & 10 & 30 & 50 & 40 & 30 & 20 \\ 25\% & 70 & 80 & 80 & 70 & 60 & 40 \\ 50\% & 232 & 172 & 124 & 108 & 87 & 75 \\ 75\% &1000 & 400 & 200 & 160 & 100 & 200 \\ 95\% &30000&3000 & 400 & 300 & 400 & 600 \\ \end{tabular} \caption{ The fluxes (J$_{23}$) corresponding to various posterior probabilities for model {\bf D}. See the text for details on the expected errors in these values.} \label{tab:modeld} \end{table} \begin{figure*} \hbox{ \epsfxsize=8.5cm \epsfbox{flux_bothdy2.ps} \hfill \epsfxsize=8.5cm \epsfbox{box_bothy2.ps} } \epsfverbosetrue \caption{The expected probability distribution of the log background flux for model {\bf D}, comparing the results from the full data set with those obtained when the column density completeness limit is raised by $\Delta\log(N)=0.5$ (dashed line, left; larger boxes, right). The box plots show median, quartiles, and 95\% limits.} \label{fig:flux_both} \end{figure*} \section{Discussion} \label{sec:discuss} \subsection{Population Parameters} \label{sec:dispop} Parameter values for the different models are given in table \ref{tab:fpars}. They are generally consistent with other estimates (Lu, Wolfe \& Turnshek 1991; Rauch et~al. 1993). Including the proximity effect increases $\gamma$ by $\sim0.2$. Although not statistically significant, the change is in the sense expected, since local depletions at the higher redshift end of each data set are removed. Figure \ref{fig:var_d}\ shows the change in population parameters as the completeness limits for the observations are increased. The most likely values (curve peaks) of both $\beta$ and $\gamma$ increase as weaker lines are excluded, although $\gamma$ decreases again for the last sample. The value of $\beta$ found here ($\sim 1.7$) is significantly different from that found by Press \& Rybicki (1993) ($\beta \sim 1.4$) using a different technique (which is insensitive to Malmquist bias and line blending). The value of $\beta$ moves still further away as the column density completeness limits are increased. This is not consistent with Malmquist bias, which would give a smaller change in $\beta$ (section~\ref{sec:malm}), but could be a result of either line blending or a population in which $\beta$ increases with column density. The latter explanation is also consistent with Cristiani et~al. (1995) who found a break in the column density distribution with $\beta$ = 1.10 for log($N$)$ < 14.0$, and $\beta$ = 1.80 above this value. Later work (Giallongo et~al. 1996, see section~\ref{sec:prevhi}) confirmed this. Recent work by Hu et~al. (1995), however, using data with better signal--to--noise and resolution, finds that the distribution of column densities is described by a single power law ($\beta\sim 1.46$) until $\hbox{$\log( N )$}\sim 12.3$, when line blending in in their sample becomes significant. It might be possible that their sample is not sufficiently large (66 lines with $\log(N)>14.5$, compared with 192 here) to detect a steeper distribution of high column density lines. The change in $\gamma$ as completeness limits are raised may reflect the decrease in line blending at higher column densities. This suggests that the value here is an over--estimate, although the shift is within the 95\% confidence interval. No estimate is significantly different from the value of 2.46 found by Press \& Rybicki (1993) (again, using a method less susceptible to blending problems). \subsection{The Proximity Effect} \subsubsection{Is the Proximity Effect Real?} \label{sec:disevid} The likelihood ratio statistic (equivalent to the `F test'), comparing model {\bf A} with any other, indicates that the null hypothesis (that the proximity effect, as described by the model here, should be disregarded) can be rejected with a confidence exceeding 99.9\%. Note that this confirmation is based on the likelihood values in table~\ref{tab:fpars}. This test is much more powerful than the K--S test (section~\ref{sec:finalq}) which was only used to see whether the models were sufficiently good for the likelihood ratio test to be used. To reiterate: if model {\bf A} and model {\bf B} are taken as competing descriptions of the population of Lyman--$\alpha$\ clouds, then the likelihood ratio test, which allows for the extra degree of freedom introduced, strongly favours a description which includes the proximity effect. The model without the proximity effect is firmly rejected. This does not imply that the interpretation of the effect (ie.\ additional ionization by background radiation) is correct, but it does indicate that the proximity effect, in the restricted, statistical sense above, is `real' (cf. R\"{o}ser 1995). If the assumptions behind this analysis are correct, in particular that the proximity effect is due to the additional ionising flux from the quasar, then the average value of the background is 100\elim{50}{30}~J$_{23}$\ (model {\bf B}). If a more flexible model for the background (two power laws) is used the flux is consistent with a value of 120\elim{110}{50}~J$_{23}$\ (model {\bf D} at $z=3.25$). \subsubsection{Systematic Errors\label{sec:syserr}} Five sources of systematic error are discussed here: Malmquist bias and line blending; reddening by damped absorption systems; increased clustering of clouds near quasars; the effect of gravitational lensing. The constraints on the background given here may be affected by Malmquist bias and line blending (sections \ref{sec:malm} and \ref{sec:dispop}). The effects of line blending will be discussed further in section~\ref{sec:gerry}, where a comparison with a different procedure suggests that it may cause us to over--estimate the flux (by perhaps $0.1$ dex). Malmquist bias is more likely to affect parameters sensitive to absolute column densities than those which rely only on relative changes in the observed population. So while this may have an effect on $\beta$, it should have much less influence on the inferred background value. Attenuation by intervening damped absorption systems will lower the apparent quasar flux and so give an estimate for the background which is too low. This is corrected in model {\bf E}, which includes adjustments for the known damped systems (section \ref{sec:dampcor}, table~\ref{tab:damp}). The change in the inferred background flux is insignificant (figure \ref{fig:evoln}, table~\ref{tab:fpars}), implying that the magnitude of the bias is less than 0.1 dex. If quasars lie in regions of increased absorption line clustering (Loeb \& Eisenstein 1995; Yurchenko, Lanzetta \& Webb 1995) then the background flux may be overestimated by up to 0.5, or even 1, dex. Gravitational lensing may change the apparent brightness of a quasar --- in general the change can make the quasar appear either brighter or fainter. Absorption line observations are made towards the brightest quasars known (to get good quality spectra). Since there are more faint quasars than bright ones this will preferentially select objects which have been brightened by lensing (see the comments on Malmquist bias in section~\ref{sec:malm}). An artificially high estimate of the quasar flux will cause us to over--estimate the background. Unfortunately, models which assess the magnitude of the increase in quasar brightness are very sensitive to the model population of lensing objects. From Pei (1995) an upper limit consistent with observations is an increase in flux of about 0.5 magnitudes, corresponding to a background estimate which is too high by a factor of 1.6 (0.2 dex). The probable effect, however, could be much smaller (Blandford \& Narayan 1992). If bright quasars are more likely to be lensed we can make a rudimentary measurement of the effect by splitting the data into two separate samples. When fitted with a constant background (model {\bf B}) the result for the five brightest objects is indeed brighter than that for the remaining six, by 0.1 dex. The errors, however, are larger (0.3 dex), making it impossible to draw any useful conclusions. The effects of Malmquist bias, line blending and damped absorption systems are unlikely to change the results here significantly. Cloud clustering and gravitational lensing could be more important --- in each case the background would be over--estimated. The magnitude of these last two biases is not certain, but cloud clustering seems more likely to be significant. \subsubsection{Is there any Evidence for Evolution?} \label{sec:noevoln} More complex models allow the background flux to vary with redshift. If the flux does evolve then these models should fit the data better. However, there is no significant change in the fit when comparing the likelihood of models {\bf C} to {\bf E} with that of {\bf B}. Nor are $\alpha_1$ or $\alpha_2$ significantly different from zero. So there is no significant evidence for a background which changes with redshift. The asymmetries in the wings of the posterior distributions of $\alpha_1$ or $\alpha_2$ for model {\bf D} (figure \ref{fig:alphas_d}) are a result of the weak constraints on upper limits (see next section). The box plots in figure \ref{fig:flux_both} illustrate the range of evolutions that are possible. \subsubsection{Upper and Lower Limits\label{sec:lims}} \begin{figure*} \hbox{ \epsfxsize=8.5cm \epsfbox{delta_z.ps} \hfill \epsfxsize=8.5cm \epsfbox{delta_d.ps} } \epsfverbosetrue \caption{On the left, absorber redshifts are plotted against $\Delta_F$. On the right $\Delta_F$ is plotted against the distance between cloud and quasar. Note that the correction for the quasar's flux, and hence the upper limit to the estimate of the background, is significant for only a small fraction of the clouds.} \label{fig:prox} \end{figure*} While there is little evidence here for evolution of the background, the upper limits to the background flux diverge more strongly than the lower limits at the lowest and highest redshifts. Also, the posterior probability of the background is extended more towards higher values. The background was measured by comparing its effect with that of the quasar. If the background were larger the quasar would have less effect and the clouds with $\Delta_F < 1$ would not need as large a correction to the observed column density for them to agree with the population as a whole. If the background was less strong then the quasars would have a stronger influence and more clouds would be affected. The upper limit to the flux depends on clouds influenced by the quasar. Figure \ref{fig:prox} shows how $\Delta_F$ changes with redshift and proximity to the background quasar. From this figure it is clear that the upper limit is dominated by only a few clouds. However, the lower limit also depends on clouds near to, but not influenced by, the quasar. This involves many more clouds. The lower limit is therefore stronger, more uniform, and less sensitive to the amount of data, than the upper limit. Other procedures for calculating the errors in the flux have assumed that the error is symmetrical (the only apparent exception is Fern\'{a}ndez--Soto et~al. (1995) who unfortunately had insufficient data to normalize the distribution). While this is acceptable for $\beta$ and $\gamma$, whose posterior probability distributions (figure~\ref{fig:beta_gamma_d}) can be well--approximated by Gaussian curves, it is clearly wrong for the background (eg. figure~\ref{fig:flux_both}), especially where there are less data (at the lowest and highest redshifts). An estimate based on the assumption that the error is normally distributed will be biased in two ways. First, since the extended upper bound to the background has been ignored, it will underestimate the `best' value. Second, since the error bounds are calculated from the curvature of the posterior distribution at its peak (ie. from the Hessian matrix) they will not take account of the extended `tails' and so will underestimate the true range of values. In addition, most earlier work has calculated errors assuming that the other population parameters are fixed at their best--fit values. This will also under--estimate the true error limits. All these biases become more significant as the amount of data decreases. The first of these biases also makes the interpretation of the box--plots (eg. figures \ref{fig:flux_both} and \ref{fig:evoln}) more difficult. For example, the curves in the left--hand plot in figure~\ref{fig:flux_both} and the data in table~\ref{tab:modeld} show that the value of the flux with highest probability at $z=2$ is $140$~J$_{23}$\ (for model {\bf D}). In contrast the box--plot on the right shows that the median probability is almost twice as large ($230$~J$_{23}$). Neither plot is `wrong': this is the consequence of asymmetric error distributions. \begin{figure*} \epsfxsize=15.cm \epsfbox{box_bothe3.ps} \epsfverbosetrue \caption{The expected probability distribution of the log background flux for models {\bf D} (left) and {\bf E} (right, including a correction for the known damped absorption systems). The box plots show median, quartiles, and 95\% limits. The shaded area covers the range of backgrounds described in Fall \& Pei (1995). The lower boundary is the expected background if all quasars are visible, the higher fluxes are possible if an increasing fraction of the quasar population is obscured at higher redshifts. The crosses and arrows mark the extent of previous measurements from high resolution spectra --- see the text for more details.} \label{fig:evoln} \end{figure*} \subsection{Comparison with Previous Estimates} \subsubsection{Earlier High--Resolution Work} \label{sec:prevhi} Fern\'{a}ndez--Soto et~al. (1995) fitted high signal--to--noise data towards three quasars. For $2 < z < 2.7$\ they estimate an ionizing background intensity of 32~J$_{23}$, with an absolute lower limit (95\% confidence) of 16~J$_{23}$\ (figure~\ref{fig:evoln}, the leftmost cross). They were unable to put any upper limit on their results. Cristiani et~al. (1995) determined a value of 50~J$_{23}$\ using a sample of five quasars with a lower column density cut--off of log($N$) = 13.3. This sample was recently extended (Giallongo 1995). They find that the ionizing background is roughly constant over the range $1.7 < z < 4.0$\ with a value of 50~J$_{23}$\ which they considered a possible lower limit (figure~\ref{fig:evoln}, the middle lower limit). While this paper was being refereed Giallongo et~al. (1996) became available, extending the work above. Using a maximum likelihood analysis with an unspecified procedure for calculating errors they give an estimate for the background of $50\pm10$~J$_{23}$. They found no evidence for evolution with redshift when using a single power law exponent. Williger et~al. (1994) used a single object, Q1033--033, which is included in this sample, to give an estimate of $10-30$~J$_{23}$\ (figure~\ref{fig:evoln}, the rightmost cross). The error limits are smaller than those found here, even though they only use a subset of this data, which suggests that they have been significantly underestimated. If the errors in Williger et~al. (1994) are indeed underestimates then these measurements are consistent with the results here. However, the best--fit values are all lower than those found here. This may be, at least partly, because of the biases discussed in section~\ref{sec:lims}. Williger et~al. (1994) used a more direct method than usual to estimate the background. This gives a useful constraint on the effect of line blending in the procedures used, and is explored in more detail below. \subsubsection{Q1033--033 and Line Blending\label{sec:gerry}} The measured value of the background, 80\elim{80}{40}~J$_{23}$\ (model {\bf D} at $z=4$), is larger than an earlier estimate using a subset of this data (Williger et~al. 1994, Q1033--033, $10-30$~J$_{23}$). As has already been argued, it is difficult to understand how a procedure using much less data could have smaller error limits than the results here, so it is likely that the error was an underestimate and that the two results are consistent. However, it is interesting to see if there is also a systematic bias in the analyses used. The correction for galactic absorption is not very large for this object (about 20\%). More importantly, the procedures used differ significantly in how they are affected by blended lines. These are a problem at the highest redshifts, where the increased Lyman--$\alpha$\ cloud population density means that it is not always possible to resolve individual clouds. Williger et~al. (1994) added additional lines ($\hbox{$\log( N )$} = 13.7$, $b$ = 30~\hbox{\rm km\thinspace s$^{-1}$}) to their $z$ = 4.26 spectra of Q1033--033 and found that between 40\% and 75\% would be missed in the line list. As the lower column density limit is raised Williger et~al. (1994) find that the observed value of $\gamma$ also increases. The resulting stronger redshift evolution would make the deficit of clouds near the quasar more significant and so give a lower estimate of the background. Although not significant at the 95\% level, there is an indication that $\gamma$ also increases with higher column density in this analysis (section \ref{sec:dispop}, figure \ref{fig:var_d}). While it is possible that $\gamma$ varies with column density the same dependence would be expected if line blending is reducing the number of smaller clouds. To understand how line blending can affect the estimates, we will now examine the two analyses in more detail. Line blending makes the detection of lines less likely. Near the quasar lines are easier to detect because the forest is more sparse. In the analysis used in this paper the appearance of these `extra' lines reduces the apparent effect of the quasar. Alternatively, one can say that away from the quasar line blending lowers $\gamma$. Both arguments describe the same process and imply that the estimated background flux is too large. In contrast, Williger et~al. (1994) take a line--list from a crowded region, which has too few weak lines and correspondingly more saturated lines, and reduce the column densities until they agree with a region closer to the quasar. Since a few saturated lines are less sensitive to the quasar's flux than a larger number of weaker lines, the effect of this flux is over--estimated (and poorly determined), making the background seem less significant and giving a final value for the background flux which is too small. This method is therefore biased in the opposite sense to ours and so the true value of the background probably lies between their estimate and ours. The comparison with Williger et~al. (1994) gives one estimate of the bias from line blending. Another can be made by raising the completeness limits of the data (section~\ref{sec:malm}). This should decrease the number of weak, blended lines, but also excludes approximately half the data. In figure \ref{fig:flux_both} the flux estimates from the full data set are shown together with those from one in which the limits have been raised by $\Delta\log(N)=0.5$. There is little change in the lowest reasonable flux, an increase in the upper limits, and in increase in the `best--fit' values. The flux for $z<3$ is almost unconstrained by the restricted sample (section~\ref{sec:lims} explains the asymmetry). An increase of 0.5 in $\log(N)$ is a substantial change in the completeness limits. That the lower limits remain constant (to within $\sim 0.1$ dex) suggests that line blending is not causing the flux to be significantly over--estimated. The increase in the upper limits is expected when the number of clouds in the sample decreases (section~\ref{sec:lims}). In summary, the total difference between our measurement and that in Williger et~al. (1994) is 0.7 dex which can be taken as an upper limit on the effect of line blending. However, a more typical value, from the constancy of the lower limits when completeness limits are raised, is probably $\sim0.1$ dex. \subsubsection{Results from Lower Resolution Spectra} Bechtold (1994) analysed lower resolution spectra towards 34 quasars using equivalent widths rather than individual column density measurements. She derived a background flux of 300~J$_{23}$\ ($1.6<z<4.1$), decreasing to 100~J$_{23}$\ when a uniform correction was applied to correct for non--systemic quasar redshifts. With low--resolution data a value of $\beta$ is used to change from a distribution of equivalent widths to column densities. If $\beta$ is decreased from 1.7 to a value closer to that found for narrower lines (see section~\ref{sec:dispop}) then the inferred background estimate could decrease further. The evolution was not well--constrained ($-7<\alpha<4$). No distinction was made between the lower and upper constraints on the flux estimate, and it is likely that the wide range of values reflects the lack of strong upper constraints which we see in our analysis. It is not clear to what extent this analysis is affected by line blending. Certainly the comments above --- that relatively more clouds will be detected near the quasar --- also apply. \subsubsection{Lower Redshift Measurements} The background intensity presented in this paper is much larger than the 8~J$_{23}$\ upper limit at $z=0$ found by Vogel et~al. (1995). Kulkarni \& Fall (1993) obtain an even lower value of 0.6\elim{1.2}{0.4}~J$_{23}$\ at $z=0.5$ by analysing the proximity effect in HST observations. However, even an unevolving flux will decrease by a factor of $\sim 50$ between $z=2$ and $0$, so such a decline is not inconsistent with the results given here. \subsection{What is the Source of the Background?} \label{sec:source} \subsubsection{Quasars} Quasars are likely to provide a significant, if not dominant, contribution to the extragalactic background. An estimate of the ionizing background can be calculated from models of the quasar population. Figure \ref{fig:evoln} shows the constraints from models {\bf D} and {\bf E} and compares them with the expected evolution of the background calculated by Fall \& Pei (1995). The background can take a range of values (the shaded region), with the lower boundary indicating the expected trend for a dust--free universe and larger values taking into account those quasars that may be hidden from our view, but which still contribute to the intergalactic ionizing flux. The hypothesis that the flux is only from visible quasars (the unobscured model in Fall \& Pei 1995) is formally rejected at over the 95\% significance level since the predicted evolution is outside the 95\% bar in the box plots at higher redshift. Although our background estimate excludes a simple quasar--dominated model based on the observed number of such objects, the analysis here may give a background flux which is biased (too large) from a combination of line blending (section~\ref{sec:gerry}) and clustering around the background quasars. From the comparison with Williger et~al. (1994), above, there is an upper limit on the correction for line blending, at the higher redshifts, of 0.7 dex. However, an analysis of the data when column density completeness limits were increased by $\Delta\log(N)=0.5$ suggests that a change in the lower limits here of $\sim 0.1$ dex is more likely. A further change of up to between 0.5 and 1 dex is possible if quasars lie in regions of increased clustering (section~\ref{sec:syserr}). These two effects imply that at the highest redshifts the flux measured here could reasonably overestimate the real value by $\sim 0.5$ dex. This could make the measurements marginally consistent with the expected flux from the observed population of quasars. There is also some uncertainty in the expected background from quasars since observations could be incomplete even at the better understood lower redshifts (eg.~Goldschmidt et~al. 1992) and while absorption in damped systems is understood in theory (Fall \& Pei 1993) its effect is uncertain (particularly because the distribution of high column density systems is poorly constrained). The highest flux model (largest population of obscured quasars) from Fall \& Pei (1995) is consistent with the measurements here (assuming that the objects used in this paper are not significantly obscured). \subsubsection{Stars} The background appears to be stronger than the integrated flux from the known quasar population. Can star formation at high redshifts explain the discrepancy? Recent results from observations of low redshift starbursts (Leitherer et~al. 1995) suggest that very few ionizing photons ($\leq 3\,$\%) escape from these systems. If high redshift starbursts are similar in their properties, then the presence of cool gas in these objects would similarly limit their contribution to the ionizing background. However, Madau \& Shull (1996) estimate that if star formation occurs in Lyman--$\alpha$\ clouds, and a significant fraction of the ionizing photons ($\sim 25\,\%$) escape, then these photons may contribute a substantial fraction of the ionizing background photons in their immediate vicinity. As an example, at $z \sim 3$\ they estimate that $J_\nu \leq 50~\hbox{J$_{23}$}$\ if star formation sets in at $z\sim3.2$. This flux would dominate the lowest (no correction for obscuration) quasar background shown in figure \ref{fig:evoln} and could be consistent with the intensity we estimate for the background at this redshift, given the possible systematic biases discussed above and in section~\ref{sec:syserr}. \section{Conclusions} \label{sec:conc} A model has been fitted to the population of Lyman--$\alpha$\ clouds. The model includes the relative effect of the ionising flux from the background and nearby quasars (section~\ref{sec:model}). The derived model parameters for the population of absorbers are generally consistent with earlier estimates. There is some evidence that $\beta$, the column density power law population exponent, increases with column density, but could also be due to line blending (section~\ref{sec:dispop}). The ionising background is estimated to be 100\elim{50}{30}~J$_{23}$\ (model {\bf B}, section~\ref{sec:resbackg}) over the range of redshifts ($2<z<4.5$) covered by the data. No strong evidence for evolution in the ionizing background is seen over this redshift range. In particular, there is no significant evidence for a decline for $z>3$ (section~\ref{sec:noevoln}). Previous results may have been biased (too low, with optimistic error limits --- section~\ref{sec:lims}). Constraints on the evolution of the background are shown in figure \ref{fig:evoln}. The estimates are not consistent with the background flux expected from the observed population of quasars (section~\ref{sec:source}). However, two effects are likely to be important. First, both line blending and increased clustering of clouds near quasars lead to the measured background being overestimated. Second, a significant fraction of the quasar population at high redshifts may be obscured. Since their contribution to the background would then be underestimated this would imply that current models of the ionizing background are too low. Both of these would bring the expected and measured fluxes into closer agreement. It is also possible that gravitational lensing makes the measurement here an overestimate of the true background. The dominant source of errors in our work is the limited number of lines near the background quasar (eg. figures \ref{fig:nu1_d}\ and \ref{fig:prox}). Systematic errors are smaller and become important only if it is necessary to make standard (unobscured quasar) models for the background consistent with the lower limits presented here. Further data will therefore make the estimate here more accurate, although observational data are limited by confusion of the most numerous lower column density systems ($\hbox{$\log( N )$}< 13.0$) so it will remain difficult to remove the bias from line blending. An improvement in the errors for the highest redshift data points, or a determination of the shape of the ionizing spectrum (e.g. from He~II/H~I estimates in Lyman--$\alpha$\ clouds) would help in discriminating between current competing models for the ionizing background. Finally, a determination of the background strength in the redshift range $0.5 < z < 2.0$ is still needed. \section{Acknowledgements} We would like to thank Yichuan Pei for stimulating discussions and for making data available to us. Tom Leonard (Dept. of Statistics, Edinburgh) gave useful comments and guidance on the statistics used in this paper. We would also like to thank an anonymous referee for helpful and constructive comments.
proofpile-arXiv_065-429
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} The study of the cosmic velocity field is a very promising and crucial area for the understanding of large--scale structure formation. Since the early work of Rubin et al. (1976) and Burstein et al. (1987) a lot of effort has been invested in the measurement of the large--scale velocity flows (see Dekel 1994 and Strauss \& Willick 1995 for recent reviews of the subject). Particularly important for the analysis of these measured velocity fields has been the development of the parameter-free POTENT method by Bertschinger \& Dekel (1989). Based on the plausible assumption of potential flow it enabled the construction and study of the full three-dimensional velocity field in a fair fraction of the local Universe out of the measurements of galaxy line-of-sight peculiar velocities. This opened up and triggered a host of studies addressing various issues and aspects of cosmic velocity flows and provided a versatile ground for testing the scenario of large--scale structure formation. The cosmic velocity field is particularly interesting because of its close and direct relation to the underlying field of mass fluctuations. Indeed, on these large scales the acceleration, and therefore the velocity, of any object is expected to have an exclusively gravitational origin so that it should be independent of its nature, whether it concerns a dark matter particle or a bright galaxy. Moreover, the linear theory of the generic gravitational instability scenario predicts that at every location in the Universe the local velocity is related to the local acceleration, and hence the local mass density fluctuation field, through the same universal function $f(\Omega)\approx\Omega^{0.6}$ (Peebles 1980). As the linear theory provides a good description on those large scales the use of this straightforward relation implies the possibility of a simple inversion of the measured velocity field into a field that is directly proportional to the field of local mass density fluctuations $\delta = \rho/\mg\rho\md -1$. Such a procedure can be used to infer the value of $\Omega$, through a comparison of the resulting field with the field of mass density fluctuations in the same region. However, the determination of this mass density fluctuation field through the measurement of the local galaxy density fluctuation field, $\delta_g$, may be contrived. The galaxy distribution may be representing a biased view of the underlying mass density fluctuation field. A common and rather simplistic assumption is that $\delta_g$ and $\delta$ are related via a linear bias factor $b$, \begin{equation} \delta_g(\vr)=b\ \delta(\vr). \label{eq:def_bias} \end{equation} However, although several physical mechanisms have been invoked to explain such a {\it linear bias model} (see e.g. Dekel \& Rees 1987), by lack of a complete and self-consistent theory of galaxy formation it should as yet only be considered as a numerical factor roughly describing the contrast of galaxy density fluctuations with respect to the mass density fluctuations. The comparison between the observed local galaxy density fluctuation field and the local cosmic velocity field, invoking equation~(\bibitem{eq:def_bias}), will therefore provide an estimate of the ratio, \begin{equation} \beta=\frac{f(\Omega)}{b}\approx \frac{\Omega^{0.6}}{b}. \end{equation} Various studies, most notably the ones based on a comparison of the galaxy density field inferred from the IRAS redshift survey of Strauss et al. (1990) and the local velocity field reconstructed by the POTENT algorithm (Bertschinger et al. 1990; Dekel, Bertschinger \& Faber 1990; see Dekel 1994), have yielded estimates of $\beta$ in the range $\beta \approx 0.5-1.2$ (see Dekel 1994, Strauss \& Willick 1995, for compilations of results). \begin{figure} \vskip 7.5 cm \special{hscale=50 vscale=50 hoffset=0 voffset=-80 psfile=sketch.ps} \caption{ The PDF of the velocity divergence as given by Eq. (\bibitem{eq:PDF_theor_exact}) showing its dependence with $\Omega$ and $\sigma$.} \label{fig:PDF_theor} \end{figure} It is then of crucial interest to find ways to disentangle the contribution of $\Omega$ and $b$ to $\beta$. A variety of methods using {\it intrinsic} properties of the large-scale velocity field have been proposed to achieve this. One such attempt is based on the reconstruction of the initial density field from the observed distribution of matter through the use of the Zel'dovich approximation. The further assumption of Gaussianity of the initial density probability distribution leads to a constraint on $\Omega$ (Nusser \& Dekel 1993). Although promising, the quantitative results of this approach may be questionable, as the Zel'dovich approximation provides non-exact results for the induced non-Gaussian properties of the density and velocity field (Bouchet et al 1992, Bernardeau 1994a). Another interesting attempt is the one by Dekel \& Rees (1994), who exploit the simple observation that voids have a maximal ``emptiness'' of $\delta=-1$. On the basis of the corresponding analysis of the measured velocity field in and around a void region these authors inferred a lower limit on $\Omega$ of about 0.3. In this paper we focus on a method to determine the value of $\Omega$ that finds its origin in a statistical analysis of the velocity field. The foundation for this method is formed by analytical work within the context of the perturbation theory for the evolution of density and velocity fluctuations. In particular, it focuses on the statistical properties of the divergence of the locally smoothed velocity field $\theta$, which is defined as \begin{equation} \theta \equiv \frac{\nabla \cdot {\bf v}}{H}, \label{eq:def_theta} \end{equation} where $\bf v$ is the time derivative of the comoving coordinate $\bf x$, and $\nabla = {\partial}/{\partial \bf x}$. The method, proposed by Bernardeau (1994a) and Bernardeau et al. (1995), exploits the relations between the lower order moments of the probability distribution function (PDF) of $\theta$, and the explicit dependence on $\Omega$ of these relations. The specific form of these relations were derived under the assumption of Gaussian initial conditions. The viability of the method was demonstrated by Bernardeau et al. (1995), who used the strong $\Omega$-dependence of the skewness factor of $\theta$, or rather the third normalized moment $T_3$, \begin{equation} T_3 \equiv \mg\theta^3\md/\mg\theta^2\md^2 \propto \Omega^{-0.6} \end{equation} to estimate successfully the density parameter in N-body simulations of structure formation. A tentative application of this method to the observed velocity field as processed by the POTENT method yielded results consistent with $\Omega=1$. Moreover, subsequent work by Bernardeau \& van de Weygaert (1996) showed that the theoretical predictions concerning the complete overall shape of the PDF of $\theta$ were valid. They demonstrated this by comparison of the analytically predicted PDF with the PDF determined from a large CDM N-body simulation with $\Omega=1$. In order to deal with the major complication of obtaining clean and straightforward numerical estimates of the statistical properties of the velocity divergence field from the discretely sampled velocity field, they developed two new techniques. These techniques exploit the minimum triangulation properties of Voronoi and Delaunay tessellations (Voronoi 1908, Delaunay 1934, see Van de Weygaert 1991, 1994 for references and applications). A Voronoi tessellation of a set of particles is a space filling network of polyhedral cells, with each cell being defined by one of the particles as its nucleus and delimiting the part of space closer to this nucleus than to any other of the particles. Closely related to the Voronoi tessellation is the Delaunay tessellation, a space filling lattice of tetrahedra (in three dimensions). Each of the tetrahedra in the Delaunay tessellation has four particles of the set as its vertices, such that the corresponding circumscribing sphere does not have any other particle inside. Through the duality relation of the Voronoi and the Delaunay tessellation it is possible to obtain one from the other. In the method based on the Voronoi tessellation -- the ``Voronoi method'' -- the velocity field is defined to be uniform within each Voronoi polyhedron, with the velocity at every location within each cell being equal to that of its nucleus. The obvious implication of such an interpolation scheme is that the only regions of space where the velocity divergence, as well as the shear and vorticity, acquires a non-zero value is in the polygonal walls that separate the cells. While the Voronoi method can be regarded as a zeroth order interpolation scheme, yielding a discontinuous velocity field, the ``Delaunay method'' can be seen as the corresponding first order scheme. Basically it constructs the velocity field within each Delaunay tetrahedron through linear interpolation between the velocities of the four defining particles. Evidently, the velocity field constructed by the Delaunay method is a field of uniform velocity field gradients within each Delaunay cell. In the following we consider the statistical properties of the local velocity divergence when it is filtered by a top-hat window function following a {\it volume weighting} prescription. In practice the local filtered divergence is computed on $50^3$ grid points in all cases. Depending on the method used to define the velocity field, the filtered divergence is given by a sum of intersections of a sphere with either planar polygonal walls or tetrahedra (each of them being multiplied by the local divergence) divided by the volume of the sphere. In such a volume weighting scheme the filtered quantities do not depend on the local number density of tracers. But of course the more numerous they are, the more accurately determined are the local divergences. For an extensive description and preliminary tests of the two techniques we refer to Bernardeau \& van de Weygaert (1996). With the intention of demonstrating the potential and practical applicability of our method, this paper presents the results of a systematic study of the $\Omega$ dependence of the moments and the PDF of the velocity divergence $\theta$ in N-body simulations of structure formation, using the numerical schemes of the Voronoi and the Delaunay method. To this end, we will first recall the relevant theoretical results on the statistical properties of $\theta$ in Section \bibitem{sec:PT}, thereby underlining the main features of importance to our study. This theoretical groundwork is followed by Section \bibitem{sec:num}, containing the presentation of the numerical results of the statistical analysis of PM N-body simulations of structure formation in $\Omega=1$ and $\Omega<1$ universes. While the first subsection, \S~\bibitem{sec:large_sample}, concerns the statistical quantities that were determined in a very large sample and thus with maximum attainable accuracy, the second subsection \S~\bibitem{sec:dilut} is devoted to the issue to what extent these results get affected when the sample is strongly diluted. The latter is particularly important as we are interested in the reliability of our method for samples with a number density comparable to that of available galaxy catalogues. Also of immediate relevance for a practical application is the question of how far the results of the statistical analysis get influenced when the velocity of the particles in the sample are known along only one direction. This issue, of crucial importance within the context of the statistical analysis of observational catalogues, is treated in a third subsection, in \S~\bibitem{sec:1D}. Finally, following the successful application of our method under the circumstances described above, we conclude with a summary and a discussion of possible complications and prospects for our statistical method to infer a bias-independent value of $\Omega$. \section{Perturbation Theory of Structure Formation and the velocity field Probability Distribution Function} \label{sec:PT} Perturbation Theory (PT) is extremely useful for the study and analytical description of the mildly non linear evolution of density and velocity fields. In particular within the context of structure forming out of Gaussian initial conditions perturbation theory has been extensively developed in a large body of work (see e.g. Bernardeau 1994a,b). In the case of these Gaussian initial conditions the complete set of moments of the smoothed velocity and density fields can be computed analytically, in particular if these fields are top-hat filtered. The corresponding PDF can be computed through re-summation of the series of moments. One of the most straightforward and useful results in the context of perturbation theory is the relation between the third moment $\mg\theta^3\md$ and the second moment $\mg\theta^2\md$ of the probability distribution function of $\theta$, \begin{equation} \mg\theta^3\md=T_3\,\mg\theta^2\md^2 = T_3\,\sigma_{\theta}^4. \end{equation} The coefficient $T_3$ depends on the cosmological parameter $\Omega$, on the shape of the power spectrum, on the geometry of the window function that has been used to filter the velocity field and even on the value of the cosmological constant $\Lambda$, although the latter is an almost negligible weak dependence. In fact, the dependence of $T_3$ on $\Omega$ is substantially stronger than the one of the equivalent coefficient for density field. For instance, for a top-hat window function and a power law initial power spectrum of index $n$, i.e. \begin{equation} P(k) \equiv \mg\delta({\bf k})^2\md \propto \mg\theta({\bf k})^2\md \propto k^{n}, \end{equation} one obtains the following expression for $T_3$, \begin{equation} T_3={-1\over \Omega^{0.6}}\left[{26\over7}-(3+n)\right]. \end{equation} As $T_3$ can be directly determined from observations, one can use its strong dependence on $\Omega$ to obtain an estimate of $\Omega$, as has been done by Bernardeau et al. (1995). More generally, Perturbation Theory enables one to infer the whole set of the cumulants $\mg\theta^p\md$ to their leading order. All of them are related to the second moment via the relation \begin{equation} \mg\theta^p\md=T_p\,\mg\theta^2\md^{p-1}, \label{eq:def_Tp} \end{equation} and as in the case of $T_3$ all the coefficients $T_p$ possess a strong dependence on the value of $\Omega$ (Bernardeau 1994b). To a good approximation, this dependence on cosmological parameters can be written as \begin{equation} T_p(\Omega,\Lambda)\approx{1\over\Omega^{(p-2)\,0.6}}T_p(\Omega=1,\Lambda=0). \end{equation} This property, given here for the moments, naturally extends itself to the shape of the complete velocity divergence PDF $p(\theta)$. This can be directly appreciated from the work by Bernardeau (1994b), who showed that the PDF can be calculated from its moments $T_p$ and the value of $\sigma_{\theta}^2$ through a Laplace transform of its generating function $\varphi_{\theta}$ \begin{equation} p(\Omega,\theta){\rm d}\theta=\int_{-{\rm i}\infty}^{+{\rm i}\infty} {{\rm d} y\over 2\pi{\rm i}\sigma_{\theta}^2}\exp\left[-{ \varphi_{\theta}(\Omega,y)\over \sigma_{\theta}^2} +{ y\theta\over \displaystyle \sigma_{\theta}^2}\right] {\rm d}\theta\,. \label{eq:PDF_theor_exact} \end{equation} The moment generating function $\varphi_{\theta}(\Omega,y)$, given by \begin{equation} \varphi_{\theta}(\Omega,y)=\sum_{p=2}^{\infty}\ -T_p(\Omega) \frac{(-y)^p}{p!}, \label{eq:gener_f} \end{equation} can be related to the spherical collapse dynamics in the cosmology under consideration (see Appendix A). Although this calculation is almost intractable in the general case, it has been shown that at least in one particular case one can evaluate this expression analytically. In the specific case of a power law density perturbation spectrum with an index $n=-1$ in combination with a top-hat smoothing, it is possible to invoke some approximations that enable the derivation of a simple analytic fit for the PDF $p(\theta)$\footnote{This fit is obtained through approximations which tend to lower the values of the moments (appendix A). The PDF (\bibitem{eq:PDF_theor}) presented here is actually more accurate if $n\approx -1.3$.} (see Appendix A). This fit is given by \begin{eqnarray} p(\theta){\rm d}\theta &=& {([2\kappa-1]/\kappa^{1/2}+ [\lambda-1]/\lambda^{1/2})^{-3/2} \over \kappa^{3/4} (2\pi)^{1/2} \sigma_{\theta}}\, \nonumber \\ & \times & \exp\left[-{ \theta^2\over 2\lambda \sigma_{\theta}^2}\right]\,{\rm d}\theta, \label{eq:PDF_theor} \end{eqnarray} with $\kappa=1+\theta^2/\left(9\lambda \Omega^{1.2}\right),$ and $\lambda=1-2\theta/\left(3\Omega^{0.6}\right).$ \begin{figure*} \vskip 16.7 cm \special{hscale=90 vscale=90 voffset=-150 hoffset=-10 psfile=pdf_large_sample.ps} \caption{The PDF of the velocity divergence for various values of $\Omega$. The dotted lines correspond to the approximate analytic fit (Eq.\bibitem{eq:PDF_theor}) and the solid lines to the theoretical predictions using (A.15) and (A.14) with $n=-0.7$ obtained for the measured values of $\sigma$ and $\Omega$. The dashed lines are the predictions for $\Omega=1$ and the same variance. The numerical estimations have been obtained using the Delaunay method.} \label{fig:PDF_large_sample} \end{figure*} \begin{figure*} \vskip 7.8 cm \special{hscale=90 vscale=90 voffset=-390 hoffset=-10 psfile=pdf_small_sample.ps} \caption{Effects of dilution for $\Omega=0.33$ (left panels) and $\Omega=1$ (right panels). The PDF-s have been obtained with only 10,000 tracers in total which gives an average of about 10 particles per cell. The squares show the results of the Voronoi method and the triangles the results of the Delaunay method. The solid lines correspond to the theoretical predictions for the variance obtained with each method and the ``right'' assumption for $\Omega$ and the dashed line with the ``wrong'' assumption for each case and method.} \label{fig:PDF_small_sample} \end{figure*} The behaviour of this function $p(\theta)$ has been illustrated in figure \bibitem{fig:PDF_theor} for various values of $\Omega$ and $\sigma_{\theta}$. Qualitatively, one can see that the dependence of the shape on $\Omega$ reveals itself in two ways: (1) the location of its cut-off at high positive values of $\theta$ and (2) the location of its peak. As for (1), the maximum value that $\theta$ can obtain is known exactly and is not dependent on the approximations that have been invoked to derive expression (\bibitem{eq:PDF_theor}) \begin{equation} \theta_{\rm max}=1.5\ \Omega^{0.6}. \end{equation} The value of $\theta_{\rm max}$ determines the location of the cut-off, and therefore the maximum expansion rate in voids. The value of 1.5 is the difference in value of the Hubble parameter in an empty, $\Omega=0$, Universe and that of an Einstein-de Sitter Universe, $\Omega=1$. Evidently, this is reflecting the fact that the interior of the deepest voids locally mimic the behaviour of an $\Omega=0$ Universe. Recall that the suggestion by Dekel \& Rees (1994) of using the maximum emptiness of voids to constrain $\Omega$ is also based on a similar feature. Also quite sensitive to the value of $\Omega$ are the position of the peak of the distribution function $p(\theta)$, i.e. the most likely value of $\theta$, and the overall shape of $p(\theta)$. Using the Edgeworth expansion (Juszkiewicz et al. 1995, Bernardeau \& Kofman 1995) one can show that the value of $\theta$ for which the distribution reaches its maximum is given by \begin{equation} \theta_{\rm peak}\approx -{T_3\over 2}\ \sigma_{\theta}={1\over \Omega^{0.6}} \sigma_{\theta}. \end{equation} In fact, a procedure exploiting this dependence of shape and peak location of $p(\theta)$ will probably yield a more robust measure of $\Omega$ than the maximum value of $\theta_{\rm max}$ as it will be less bothered by the noise in the tails. \section{The $\Omega$ dependence in numerical simulations} \label{sec:num} By means of numerical simulations we have investigated the discussed dependence of the PDF of $\theta$. These N-body simulations use a Particle-Mesh (PM) code (Moutarde et al. 1991) with a $256^3$ grid to follow the evolution of a system of $256^3$ particles. For our project we used two simulations, one with $\Omega$ having a value of $\Omega=1$ and the second one of $\Omega < 1$. By analyzing the latter at different time-steps we explore situations for different values of $\Omega$. The particle distribution in the two simulations corresponds to a density and velocity fluctuation field with a $P(k)\propto k^{-1}$ spectrum. As can be seen in Table~\bibitem{tab:cumulant}, the variances $\sigma_{\theta}$ do not differ significantly for the different values of $\Omega$ for a given filtering radius. The fact that the values of the variance are comparable simplifies a comparison of the PDF substantially, which makes the interpretation in terms of the intrinsic $\Omega$ dependence more straightforward. \subsection{Measurements with a large number of tracers} \label{sec:large_sample} The first step of our analysis concerns an exploration of the velocity field using a large number of tracers. For this study the number of selected particles in each simulation is about 70,000, which for a cell radius of about 6\% of the box size leads to a mean number of 67 particles per cell. The selection procedure used here is deliberately biased towards low-density regions by inducing it to retain a uniform density of particles all over the simulation box. Except for its goal of achieving a better velocity field coverage of low-density regions such a selection bias is not expected to influence the velocity field analysis. The methods that we use to analyze the simulations are exactly the same ones as described by Bernardeau \& van de Weygaert (1996). In fact, at this stage we only used the Delaunay method to calculate numerically the shape of the PDF of the velocity divergence. The results for the case of a large number of velocity field tracers are shown in Fig.~\bibitem{fig:PDF_large_sample}. The results are in good agreement with the theoretical predictions for the values of $T_3$ and $T_4$ (see Table~\bibitem{tab:cumulant}), as well as with the theoretical shape of the PDF (Fig.~\bibitem{fig:PDF_theor}). It is in particular worth noting that the specific features expected from equation~(\bibitem{eq:PDF_theor}) are indeed confirmed by the numerical results. Notably, the locations of the cut-off, which are very sensitive to rare event discrepancies, are well reproduced (solid and dotted lines). Moreover, as can be observed from the insets, also the position and shape of the peak have been reproduced very well (solid lines), providing a strong discriminatory tool between different values of $\Omega$. Within the context of these observations, we should issue a few side remarks. Although the shape (\bibitem{eq:PDF_theor}) is very attractive because it is a close analytic form, one should have in mind that it is only approximate. Indeed it is derived from an approximate expression for the cumulant generating function. The differences do not reveal for the overall shape (the logarithmic plots) but are significant for the shape of the peaks. For calculating the theoretical predictions we are then forced to use a more accurate description of the cumulants. To achieve this we use the relations (A.13, A.14) with $n=-0.7$ (the expression \bibitem{eq:PDF_theor} corresponds to $n=-1$) in the integral (A.15) which is then computed numerically. It is still an approximate expression, but it yields the correct value for $T_3$ and a very good approximation for the higher order cumulants. We should emphasize that this slight modification is only instrumental in obtaining the correct shape of the PDF around its maximum, which is indeed almost entirely determined by the values of the low order moments. This may be understood for example from the properties of the Edgeworth expansion, for which we refer to Juszkiewicz et al. 1995 and Bernardeau \& Kofman 1995. \subsection{The effects of dilution} \label{sec:dilut} In order to check the robustness of the results when only a limited number of tracers for the velocity field is available, we performed numerical experiments where only 10,000 particles are used to trace the velocity field. The selection of the sample points in this diluted sample is completely random and does not invoke the specific biased selection procedure that was used in the case described in the former subsection. For this case of diluted samples, we used both the Delaunay and the Voronoi methods for analysis. Figure~\bibitem{fig:PDF_small_sample} shows the PDFs obtained with both methods, for an $\Omega=1$ simulation and for an $\Omega=0.33$ simulation. Both the Voronoi and the Delaunay methods appear to yield numerical results that are in reasonably good agreement with the predicted PDF. Particularly encouraging is the result born out by the insets, namely the fact that the shape of the peak can still be used as a strong discriminatory tool between different values of $\Omega$. Investigating Fig.~\bibitem{fig:PDF_small_sample} in somewhat more detail, we can observe that the results obtained by the two methods are affected somewhat differently by the dilution procedure. The PDFs obtained with the Voronoi method generally possess a less sharply defined tail at the side of the high $\theta$ values. This effect appears to be stronger for the low $\Omega$ case. This behaviour may originate in the fact that the divergence is localized to a limited part of space, non-zero values being confined to the walls of the tessellation. In such a situation Poisson like errors in the measurements are expected to become particularly prominent in the heavily diluted areas of the void regions. An additional difference is a slight underestimation of the values of the $T_p$ coefficients by the Voronoi method (see Table~\bibitem{tab:cumulant}). Unfortunately, we do not see a possibility to correct for such effects. On the other hand, the Delaunay method seems to be more robust against such effects. However, it tends to underestimate the value of the variance and the higher order moments. The latter is probably a consequence of the fact that the effective filtering radius tends to be larger. \subsection{The effects of reducing the information to one velocity component} \label{sec:1D} A major complication in the analysis of velocity fields under practical circumstances is the fact that only the velocity component along the line of sight can be measured. This therefore forms the second issue that we address in this paper. As yet we restrict ourselves to an artificial situation with ideal measurements, not yet to investigate the determination of the statistical quantities associated with the velocity field under realistic circumstances. To simplify furthermore our investigations we assume that we can use the approximation of an infinitely remote observer so that the radial velocity can be identified with one velocity component, namely the $x$-direction in the following. More specifically, we address the effects of the reduction of information concerning the velocity field traced by a diluted sample of points. In principle, the fact that the velocity field is only known along one direction should not pose any problem. In the usual structure formation scenarios based on gravitational instability the large scale velocity field is expected to be non-rotational, implying it to be a potential flow and therefore the gradient of a potential that can be inferred from the measurement of only one component of the velocity (Bertschinger et al. 1990, Dekel et al. 1990). \begin{figure*} \vskip 7.8 cm \special{hscale=90 vscale=90 voffset=-390 hoffset=-10 psfile=pdf_1D_sample.ps} \caption{Effects of reduction of the information on the velocity field to only one component for $\Omega=0.33$ (left panel) and $\Omega=1$ (right panel). We used the estimator (\bibitem{eq:theta_w_estim}) with $\epsilon=0.1$ to determine numerically the local divergences. The solid lines correspond to the model (\bibitem{eq:PDF_1D_estim}) obtained from (\bibitem{eq:theta_estim_model}) with the parameters $\mu$ and $\sigma_e$ given in Table~\bibitem{tab:par1D}. The dashed lines are the resulting shapes of the PDF-s when one uses the same parameters but the ``wrong'' value for $\Omega$ ($1$ instead of $0.33$ in the left panel, $0.33$ instead of $1$ in the right one).} \label{fig:PDF_1D} \end{figure*} \begin{figure*} \vskip 11.5 cm \special{hscale=90 vscale=90 voffset=-300 hoffset=-30 psfile=scatter.ps} \caption{Scatter plots that show the differences of the various estimators of the local divergence measured at 500 different random locations. } \label{fig:scatt_div} \end{figure*} In the Voronoi method non-zero values of the divergence $\theta$ are restricted to the walls of the Voronoi tessellation, where the local divergence is given by \begin{equation} \theta_{\rm wall}=\vn\,.\,\Delta \vv, \label{eq:theta_w_exact} \end{equation} with $\vn$ being the normed vector orthogonal to the wall and $\Delta \vv$ the difference of the velocities on the opposite sides of the Voronoi wall. The expression for the local vorticity is given by \begin{equation} {\bf \omega}} %\def\vomega{{\vec \omega}_{\rm wall}=\vn\,\times\,\Delta\vv. \label{eq:vortic_w_exact} \end{equation} Assuming potential flow, and hence a zero value of ${\bf \omega}} %\def\vomega{{\vec \omega}$, this implies relations between the various components of $\Delta \vv$ and the following expressions for $\theta_{\rm wall}$: \begin{equation} \theta_{\rm wall}={\Delta v_x\over n_x}={\Delta v_y\over n_y}= {\Delta v_z\over n_z}. \end{equation} In practice, this introduces the numerically unstable operation of dividing by one component $\vn$ as it can be arbitrarily close to zero. It may therefore be more reasonable to try to estimate the value of $\theta_{\rm wall}$ using the stable, but ad-hoc, prescription of \begin{equation} \theta_{\rm wall}^{\rm estim.}= \Delta v_x { n_x\over n_x^2+\epsilon}, \label{eq:theta_w_estim} \end{equation} where $\epsilon$ is a small parameter of the order of $\epsilon \approx 0.1$. Note that such a prescription is not self-consistent in reproducing the full 3D velocity field.\footnote{This can be readily appreciated from the fact that the normal $\vn$ to a wall in the Voronoi tessellation is proportional to the vector $\Delta\vr$ between the two points on each side of the wall. Thus, if it were possible to build a consistent velocity filed from the constraints (17) it would imply $\Delta\vv\propto \Delta\vr$ yielding not only a vanishing vorticity but also a vanishing shear.}Indeed, depending on the way one goes from cell A to cell B -- and there are infinitely many ways to go from A to B, even while they are direct neighbours -- one will not necessarily find the same value for $\theta_{\rm wall}^{\rm estim.}$ using the equivalent prescriptions for $\Delta v_y$ or $\Delta v_z$. The method that we adopt here will therefore certainly not be the most accurate method. However, we may expect it to be a reasonable approximation, and will therefore use it here to illustrate the properties in which we are interested. Figure~\bibitem{fig:PDF_1D} displays the results obtained with (\bibitem{eq:theta_w_estim}) for $\epsilon=0.1$. Evidently, the results get affected to quite some extent by the transition from (\bibitem{eq:theta_w_exact}) to (\bibitem{eq:theta_w_estim}). This leads to the key question whether it is still feasible to reliably recover the statistical information on $\theta$ or not. As can be observed from the scatter plots in Fig.~\bibitem{fig:scatt_div}, the change from a situation in which one has knowledge of the full velocity to one where this has been limited to only one component thereof introduces a large scatter. However, we found that it is possible to define a meaningful representation of the scatter plots in terms of the following empirical description: \begin{equation} \theta_{\rm estim.}=\mu\ (\theta+e)\,. \label{eq:theta_estim_model} \end{equation} Within this expression the coefficient $\mu$ is a constant with some fixed value. The scatter is represented by the quantity $e$, a Gaussian random variable whose value is {\em independent} of $\theta$ and which has a vanishing mean. Using the measured values of the variance and the skewness of the distribution it is possible to estimate the value of $\mu$ and the value of the rms fluctuation $\sigma_e$ of $e$. From (\bibitem{eq:theta_estim_model}) one can readily infer that \begin{eqnarray} \sigma_{\rm estim.}&=&\mu\ \sqrt{\sigma_{\theta}^2+\sigma_e^2} \nonumber \\ T_{3\rm\ estim.}&=&{\sigma_{\theta}^4\over \mu\,(\sigma_{\theta}^2+\sigma_e^2)^2}\ T_3 \end{eqnarray} where $\sigma_{\theta}$ is the exact rms fluctuations of $\theta$ and $T_3$ its third cumulant, while $\sigma_{\rm estim.}$ and $T_{3\rm\ estim.}$ are the corresponding estimated values. By solving this set of equations one can find the values of $\mu$ and $\sigma_e$. Their values for various $\Omega$ and $\epsilon$ are listed in Table~\bibitem{tab:par1D}. Moreover, significant within the context of the ultimate goal of developing an unbiased estimator of $\Omega$, is that these parameters were found to be almost independent of the value of $\Omega$. More specifically, it turns out that the value of $\mu$ only depends on the adopted value of $\epsilon$, whereas $\sigma_e$ is independent of $\Omega$ and only marginally dependent on $\epsilon$. This can clearly be appreciated from the bottom right-hand panel of Fig.~\bibitem{fig:scatt_div}, which demonstrates that the two estimations of $\theta$, one based on $\epsilon=0.1$ and the other on $\epsilon=0.2$, are basically proportional to each other. It may therefore be argued that it is quite natural to expect that the noise $e$ introduced by this method is somehow intrinsic to the distribution. In Appendix B we describe an extremely simple model based on the assumption that the relative velocity of two particles is proportional to their relative position. This allows us to compute analytically the parameters $\mu$ and $\sigma_e$ entailed by the use of the numerical scheme (\bibitem{eq:theta_w_estim}). One can show that they are both independent on $\Omega$, and only depend on $\epsilon$ and $\sigma_{\theta}$, with $\sigma_e < \sigma_{\theta}$. These analytical predictions are listed in Table~\bibitem{tab:par1D_anal}. They appear to be fairly close to their numerical measurements. The discrepancy between the analytical and numerical estimations of these parameters is due to the loss of information associated with the projection of the velocity from three to one dimensions. Although this is not fairly represented by our model, the discrepancy seems to be quite small. \begin{table*} \begin{minipage}{140mm} \caption{Cumulants from the Perturbation Theory and as estimated by the various numerical methods. \hfill} \label{tab:cumulant} \halign{\quad\hfil#\hfil\quad& \quad\hfil$#$\hfil\quad&\quad\hfil$#$\hfil\quad&\quad\hfil$#$\hfil\quad&\quad\hfil$#$\hfil\quad&\quad\hfil$#$\hfil\quad\cr \noalign{\hrule} \noalign{\medskip} \# tracers per cell & {\rm cumulants} & \Omega=0.2 & \Omega=0.33 & \Omega=0.4 & \Omega=1 \cr \noalign{\medskip} \noalign{\hrule} \noalign{\medskip} $\infty$ & T_3^{\rm PT} & -4.5 & -3.33 & -2.97 & -1.71 \cr Perturbation Theory & T_4^{\rm PT} & 31.41 & 17.22 & 13.67 & 4.55 \cr \noalign{\medskip} \noalign{\hrule} \noalign{\medskip} $\approx 65$& \sigma^{\rm Del} & 0.37\pm0.01& 0.41\pm0.01& 0.42\pm0.01&0.38\pm0.005\cr Delaunay Method & T_3^{\rm Del} & -4.52\pm0.2& -3.3\pm0.2 & -2.9\pm0.2 &-1.54\pm0.1\cr & T_4^{\rm Del} & 31.5\pm2.5 & 17.3\pm3 &12.8\pm2.5 &3.1\pm0.7\cr \noalign{\medskip} \noalign{\hrule} \noalign{\medskip} $\approx 10$& \sigma^{\rm Del} & - & 0.36\pm0.01& - & 0.35\pm0.005\cr Delaunay Method & T_3^{\rm Del} & - & -3.8\pm0.2 & - &-1.84\pm0.1\cr & T_4^{\rm Del} & - & 24.7\pm3 & - &5.4\pm1\cr \noalign{\medskip} \noalign{\hrule} \noalign{\medskip} $\approx 10$& \sigma^{\rm Vor} & - & 0.40\pm0.01& - & 0.38\pm0.003\cr Voronoi Method & T_3^{\rm Vor} & - & -3.2\pm0.2 & - & -1.51\pm0.04\cr & T_4^{\rm Vor} & - & 19.4\pm5.5 & - & 3.62\pm1.5\cr \noalign{\medskip} \noalign{\hrule} \noalign{\medskip} $\approx 10$& \sigma^{\rm Vor} & - & 0.35\pm0.006& - & 0.34\pm0.009\cr one velocity & T_3^{\rm Vor} & - & -2.40\pm0.3& - & -1.14\pm0.07\cr component, $\epsilon=0.1$ & T_4^{\rm Vor} & - & 11.9\pm4 & - & 2.9\pm2.3\cr \noalign{\medskip} \noalign{\hrule}} \end{minipage} \end{table*} \begin{table} \caption{Numerical values of the parameters $\mu$ (bias) and $\sigma_e$ (noise) introduced by the calculation (\bibitem{eq:theta_w_estim}) of the velocity divergence.} \label{tab:par1D} \centering \begin{tabular}{lcc} \hline \ $\Omega=0.33$\ &\ $\epsilon=0.1$ &\ $\epsilon=0.2$\ \\ \hline \ $\mu$ & 0.73 & 0.56 \\ \ $\sigma_e$ & 0.25 & 0.27 \\ \hline \ $\Omega=1$\ &\ $\epsilon=0.1$ &\ $\epsilon=0.2$\ \\ \hline \ $\mu$ & 0.75 & 0.57 \\ \ $\sigma_e$ & 0.24 & 0.26 \\ \hline \end{tabular} \end{table} On the basis of the simple model described in Eq.~(\bibitem{eq:theta_estim_model}) it is possible to reconstruct the corresponding shape of the distribution of $\theta_{\rm estim}$, \begin{equation} p_{\rm estim.}(\theta_{\rm estim.})= \int_{-\infty}^{+\infty} p({\theta_{\rm estim.} \over \mu}-e)\ {\exp(-e^2/2/\sigma_e^2)\over (2\,\pi)^{1/2}}\,{{\rm d} e\over\sigma_e} \label{eq:PDF_1D_estim} \end{equation} where $p(\theta)$ is given by Eq.~(\bibitem{eq:PDF_theor}). A comparison of the resulting distribution (\bibitem{eq:PDF_1D_estim}) with the measured histograms is shown in Fig.~\bibitem{fig:PDF_1D}. The agreement appears to be quite good, rendering this phenomenological description a quite valuable one. \begin{table} \caption{Analytical estimations of the parameters $\mu$ and $\sigma_e$ (see Appendix B).} \label{tab:par1D_anal} \centering \begin{tabular}{lcc} \hline \ $\Omega=0.33$\ &\ $\epsilon=0.1$ &\ $\epsilon=0.2$\ \\ \hline \ $\mu(\epsilon)$ & 0.60 & 0.49 \\ \ $\sigma_e(\epsilon,\sigma_{\theta})\ $ & 0.20 & 0.23 \\ \hline \ $\Omega=1$\ &\ $\epsilon=0.1$ &\ $\epsilon=0.2$\ \\ \hline \ $\mu(\epsilon)$ & 0.60 & 0.49 \\ \ $\sigma_e(\epsilon,\sigma_{\theta})\ $ & 0.19 & 0.22 \\ \hline \end{tabular} \end{table} Evidently, we are still able to clearly distinguish between the scenario with a high value of $\Omega$ and the ones with a low value, although the distinction is not as clear as in the previous cases based on ideal sampling circumstances. Looking into some detail, we come to the conclusion that the signal has been diluted somehow by the noise $e$, and that there is a competition between the true rms of the divergence $\sigma_{\theta}$ and the value of $\sigma_e$. In this respect it is also important to note that it is crucial for a successful determination of $\Omega$ to have good estimates of both $\mu$ and $\sigma_e$. Once these parameters have been determined, and as long as $\sigma_{\theta} > \sigma_e$, we can see no further problem in distinguishing between the different cosmological scenarios. Fig.~\bibitem{fig:PDF_1D} gives an idea of the magnitude of the discrepancy that one gets when one assumes a wrong value of $\Omega$. In the left panel this concerns the case wherein one take a value of $\Omega=1$ instead of the actual value of $\Omega=0.33$, while in the right-hand panel it concerns the reverse situation. At this point it is worthwhile to stress once more that so far we only tested the method for {\em ideal} measurements of the line-of-sight velocities. Noise in the measured values of these velocities was not yet taken into account. Moreover, besides the fact that this noise has quite a large value, an evaluation of its influence is substantially complicated as it not only concerns random measurement errors but also contains contributions from a plethora of, partially un-understood, systematic effects. In a work in preparation we will attempt to develop specific techniques for estimating the velocity divergence PDF from such noisy line-of-sight measurements. \section{Discussion and Conclusions} \label{sec:conclusion} We have tested and confirmed the validity of the strong $\Omega$ dependence of the Probability Density Function (PDF) of the velocity divergence that had been predicted on the basis of analytical Perturbation Theory calculations. These tests are based on a numerical analysis of N-body simulations of structure formation in a Universe with a Gaussian initial density and velocity fluctuation field. On the basis of this verification we may conclude that the analytical predictions of Perturbation Theory yield very accurate results for a wide range of cosmological models. The main practical implication of our work is the basis it offers for a potentially very valuable and promising estimator of the value of $\Omega$, an estimator independent of a possible bias between the distribution of galaxies and the underlying matter distribution. The successful tests presented in this work demonstrate the validity of the equations of Perturbation Theory that form a basis for the estimates of $\Omega$ which are based on the statistical properties of the velocity divergence field. Their unbiased nature finds its origin in the fact that the relations between the various statistical moments do not contain any explicit dependence on a bias between the galaxy and the matter distribution. Moreover, the estimated values of $\Omega$ are even more direct and straightforward to interpret as the relevant statistical relations only involve a very weak dependence on the cosmological constant $\Lambda$ is expected to be very weak (Bernardeau 1994a, b). Not only relations between the statistical moments of the $\theta$ distribution, also the shape and general functional behaviour provide a useful indicator for the value of $\Omega$. When we focus on the details of this functional behaviour of the velocity divergence distribution function -- illustrated in Fig.~\bibitem{fig:PDF_large_sample} and \bibitem{fig:PDF_small_sample}) -- we can draw a few conclusions with regards the practical feasibility of obtaining reliable estimates from the shape of $p(\theta)$. Both location and shape of the peak of this distribution appear to be robust indicators of the value of $\Omega$. On the other hand, the location of the maximum of the divergence $\theta$ -- i.e. the cutoff value of $p(\theta)$ -- appears to be much more sensitive to a poor sampling. This may make it harder for it to provide reliable estimates of $\Omega$ from currently available observational catalogues (see Fig.~\bibitem{fig:PDF_small_sample}). However, it is all the more encouraging that even on the basis of the cutoff value we obtained a reasonable agreement between theoretical predictions and numerical measurements. Finally, we addressed one further crucial issue towards an application of our estimation procedures to real data sets. This concerns the problem of not being able to obtain directly the full three-dimensional velocity field. Instead, the velocity of a galaxy can only be measured along the line-of-sight. In a preliminary attempt to study the consequences of this fact for the feasibility of our method, we introduced a partially empirically defined extension of our method. Despite of the extreme crudeness and rather ad-hoc nature of this algorithm to reconstruct the full velocity field, it is quite encouraging that we are able to distinguish between the velocity PDF obtained in a flat Universe and that obtained in an open Universe. The major obstacle towards a successful application of our methods therefore appears to be the one of noisy data sets and systematic sampling errors. We have not yet dealt with these problems, deferring them to a forthcoming paper. An additional and useful application of our numerical work involves a test for structure indeed having emerged through the process of gravitational growth of an initially Gaussian random density and velocity field. Having shown Perturbation Theory to be valid, we can exploit its prediction that the PDF of $\theta$ is only dependent on a few parameters, in particular $\sigma_{\theta}$ and $\Omega$. If no values of $\sigma_{\theta}$ and $\Omega$ can be found to produce an acceptable fit to the observed velocity field, this will force us to conclude that it is unlikely that the structure developed as described within the standard framework of gravitational instability and Gaussian initial conditions. In this context it is interesting to point out that a negative skewness has been observed in the currently available datasets (Bernardeau et al. 1995), which is an indication in favour of standard scenarios. Summarizing, we may conclude that the combined machinery of the analytical perturbation theory results and the developed numerical methods and their application on the intrinsic statistical properties of the velocity field provides us with a reliable new estimator of the cosmological density parameter $\Omega$. This estimator is all the more useful as it is one of the very few which will yield values of $\Omega$ completely independent of galaxy-density field biases and almost independent of the value of $\Lambda$. \section*{Acknowledgments} F. Bernardeau would like to thank IAP, where a large part of the work has been completed, for its warm hospitality. We would like to thank A. Dekel for encouraging comments and discussions. FB and RvdW are grateful for the hospitality of the Hebrew University of Jerusalem, where the last part of this contribution was finished. R. van de Weygaert is supported by a fellowship of the Royal Netherlands Academy of Arts and Sciences. Part of this work was done while EH was at the Institut d'Astrophysique de Paris (CNRS), supported by the Minist\`ere de la Recherche et de la Technologie. Additional partial support to EH was provided by the Danish National Research Foundation through the establishment of the Theoretical Astrophysics Center. The computational means were made available to us thanks to the scientific council of the Institut du D\'eveloppement et des Resources en Informatique Scientifique (IDRIS).
proofpile-arXiv_065-430
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} Elimination of fermionic degrees of freedom is very desirable in many problems of quantum field theory and elementary particle physics. First, in lattice gauge theories integration with respect to Grassmannian variables leads to serious complications of numerical simulations. The second, appearance of fermionic variables in functional integrals hampers the application of stationary phase method. As a consequence, one cannot also apply quasiclassical expansions to evaluation of functional integrals in theories with fermions except some especially simple cases. Indeed, typical Green function can be written as \begin{equation} G(x_1, ..., x_n) = \int DA D\Psi D\bar{\Psi} e^{\frac{i}{\hbar} (S_{YM}(A) + S_{ferm}(\bar{\Psi}, \Psi, A))} {\cal O}_1(x_1) ...{\cal O}_n(x_n) \label{1} \end{equation} \noindent where ${\cal O}_i$ are some operators whereas $S_{YM}$ and $S_{ferm}$ are Yang-Mills and fermionic actions respectively. Quasiclassical approximation is defined, up to some subtleties, by stationary point equations \begin{equation} \frac{\delta S_{YM}}{\delta A}=[\mbox{the source of YM field}] \label{2} \end{equation} But what must be written in R.H.S. of eq. (\ref 2) ? Of course, one cannot put \begin{equation} [\mbox{the source of YM field}]=\frac{\delta S_{ferm}}{\delta A} \label 3 \end{equation} \noindent because $S_{ferm}$ depends on Grassmanian variables. Moreover, without preliminary exception of fermionic variables one cannot write in R.H.S. of (\ref 2) nothing except zero. But this means that in zero approximation YM field can be considered as free. It seems inappropriate in all cases in which interaction is strong. So for application of quasiclassical methods as well as for facilitation of numerical simulations on the lattice fermionic variables in functional integrals of the type (\ref 1) must be integrated out and result must be represented as functional integral with respect to only bosonic variables. In other words, the theory must be bosonized. The problem of bosonization of fermionic theories has a long history. Most likely, the first example of bosonisation of fermionic theory was given by Schwinger in his famous paper \cite{1} concerning full solution of massless $\mbox{QED}_2$. Then the problem of bosonisation was investigated by many authors, but more or less exhaustive solution was obtained only in two dimensional case and in some three dimensional models (see, for instance, papers \cite{2} and references therein). In realistic four dimensional case only partial success was achieved (see, for instance, \cite 3). In fact in all proposed bosonization schemes in four dimensions it is necessary to evaluate (exactly or in some approximation) fermionic determinant -- but it is just the main problem that must be solved by means of bosonization. To author's knowledge, the only exceptions are recent papers by Lusher \cite 4 and Slavnov \cite 5 (see also \cite 6). In Ref. \cite 4 fermionic determinant on the finite lattice is represented as infinite some of bosonic determinants. In Ref. \cite 5 fermionic determinant in $D$ dimensions is expressed via bosonic one in $D+1$ dimensions. These approaches seem useful in lattice theories but they cannot be applied to investigation of quasiclassical approximation. So hitherto no quite satisfactory representation for fermionic determinant in terms of bosonic fields is known, and at present paper we will develop another approach to bosonization. Namely, we will derive pure bosonic worldline path integral representation for fermionic determinants, Green functions and Wilson loops. Worldline approach to quantum field theory also has very long history. It was originated many years ago in classical works by Feynman \cite 7 and Schwinger \cite 8 . The main idea of this approach is to represent fermionic determinants and fermionic Green functions as functional integral over trajectory of a single relativistic particle. Let us consider, for instance, fermionic determinant for $SU(N)$ Yang-Mills theory in Euclidean space: \begin{equation} D \equiv \det (i\hat{\nabla} +im) =\det (i\hat{\nabla} +im)\gamma^5 \label 4 \end{equation} \noindent where $\hat{\nabla}=\gamma^{\mu} \nabla_{\mu} =\gamma^{\mu} (\partial_{\mu} - iA_{\mu}), \ \ (\gamma^{\mu})^{\dag}=\gamma^{\mu}, \ \ \{\gamma^{\mu},\gamma^{\nu}\}=2\delta^{\mu \nu}, \ \ iA_{\mu}(x)\in su(N)$. One can write: \begin{equation} \ln D =\frac{1}{2} \ln \det[(i\hat{\nabla} +im)\gamma^5]^2= \frac{1}{2} \ln {\det} (-\nabla_{\mu}\nabla_{\mu} + \sigma^{\mu\nu}F_{\mu\nu} +m^2) \label 5 \end{equation} \noindent where \begin{equation} F_{\mu\nu}=\partial_{\mu}A_{\nu}-\partial_{\nu}A_{\mu}- i[A_{\mu},A_{\nu}], \label 6 \end{equation} \begin{equation} \sigma^{\mu\nu}=\frac{i}{4}[\gamma^{\mu},\gamma^{\nu}] \label 7 \end{equation} Further, eq. (\ref 5) can be written, up to inessential constant, as \begin{eqnarray} \ln D&=& \frac{1}{2} \mbox{tr}\: \ln (-\nabla_{\mu}\nabla_{\mu} + \sigma^{\mu\nu}F_{\mu\nu} +m^2)\nonumber\\ &=&-\frac{1}{2} \int_0^{\infty}\frac{dT}{T} e^{-m^2T} \mbox{tr}\: e^{-T( -\nabla_{\mu}\nabla_{\mu} + \sigma^{\mu\nu}F_{\mu\nu})} \label 8 \end{eqnarray} The integral (\ref 8) is divergent at $T=0$ and must be regularized. One can use, for instance, $\zeta$-function regularization: \begin{equation} \ln D =-\frac{1}{2} \frac{d}{ds} \frac{1}{\Gamma(s)} \int_0^{\infty} dT \ T^{s-1} e^{-m^2 T} \mbox{tr}\: e^{-T( -\nabla_{\mu}\nabla_{\mu} + \sigma^{\mu\nu}F_{\mu\nu})} \raisebox{-9pt}{\rule{0.4pt}{20pt}${\scriptstyle s=0}$} \label 9 \end{equation} \noindent However, in what follows we shall write, for short, formal expression (\ref 8) for $\ln D$ bearing in mind any suitable regularization. Further, trace in eq.(\ref 8) can be represented as functional integral: \begin{eqnarray} && \mbox{Tr} e^{-T( -\nabla_{\mu}\nabla_{\mu} + \sigma^{\mu\nu}F_{\mu\nu})}\nonumber\\ &=& \int _{PBC} Dq\: \mbox{tr P}\exp \left\{ -\int_0^1 dt \left(\frac{\dot{q}^2}{4T} -i\dot{q}^{\mu}A_{\mu}(q) +T \sigma^{\mu\nu}F_{\mu\nu}(q) \right) \right\} \label{10} \end{eqnarray} \noindent Here PBC means 'periodic boundary conditions'. Path ordering in (\ref{10}) corresponds to both spin and colour matrix structures. Let us suppose for a moment that spin and colour degrees of freedom are absent in (\ref{10}). (This is just the case of scalar electrodynamics). Then, integrating in (\ref 1) over fermionic fields and using for arising fermionic determinants and fermionic Green functions formulae of the type (\ref 8), (\ref{10}), one obtains formulation of quantum field theory in terms of particles interacting with gauge field $A$. This is just result of classical work bu Feynman \cite 7. In realistic models, however, one must take into account spin and colour degrees of freedom. So, to develop worldline formulation of quantum field theory, it is necessary to represent the path ordered exponent in (\ref{10}) as functional integral. In the case of QED this was done by Fradkin \cite 9 in terms of fermionic path integral. This representation and their modifications were successfully used, in particular, for construction of derivative expansion in QED \cite{10} and for investigation of complicated Feynman diagrams \cite{11}. It is also intensively used for investigation of hidden supersymmetry in fermionic theories (see, for example, \cite{12} and references therein). Recently, D'Hoker and Gagne derived fermionic path integral representation for fermionic determinants for particles coupled with arbitrary tensor field \cite{13}. In papers \cite{8}-\cite{13} fermionic path integral representations were derived and used only for spin degrees of freedom. For colour P-exponent in (\ref{10}) fermionic path integral representation also can be derived (see, for example, \cite{14}) though it seems not so elegant as one for spin P-exponent. Apropos, works \cite{14} were, to author's knowledge, the first attempts to obtain non-perturbative information about QCD in worldline formalism. But, as we already stated at the beginning of the present paper, there exist important problems for which pure bosonic worldline path integral representation for fermionic determinants and Green functions is very desirable. We see that for solution of the latter problem it is sufficient to derive bosonic worldline path integral representation for the trace of path ordered exponent \begin{equation} Z=\mbox{tr P} e^{i\int_0^1 dt B(t)} \label{11} \end{equation} \noindent with matrix $B(t) \in GL(N),\; B(0)=B(1)$. Indeed, substituting such representation with \begin{equation} B(t)=\dot{q}^{\mu}A_{\mu}(q(t)) - \sigma^{\mu\nu}F_{\mu\nu}(q(t)) \label{12} \end{equation} \noindent in (\ref{10}), one obtains desired representation for fermionic determinant. As we will show later, bosonic path integral representation for $Z$ allows also to obtain bosonic worldline path integral representation for fermionic Green functions. Different bosonic path integral representations for $Z$ in the case $B(t) \in SU(2)$ were proposed by several authors (see \cite{15}-\cite{18}). For more general case $B(t) \in su(N)$ the an analogous representation was pointed out in \cite{19}. The typical result for $B \in SU(2)$ is \begin{equation} Z=\int DS(t) \exp \left\{ \frac{i}{2} \int _0^1 dt \left[ \mbox{tr} \sigma^3 \left( S(t)B(t)S^{\dag}(t) +iS(t)\dot{S}^{\dag}(t) \right) \right] \right\} \label{13} \end{equation} The integration in (\ref{13}) is carried out over all trajectories in the group $SU(2), \; S \in SU(2), \; \sigma^3$ is a Pauli matrix. The last result has two disadvantages. First, in any parametrization of $SU(2)$ the "action" in (\ref{13}) is rather complicated non-polinomial function. This hampers the usage of (\ref{13}) in practice. The second, it appears that integral (\ref{13}) is ill-defined and needs insertion of some regulators in the "action" (see the discussion of this point in \cite{16,19} and in more recent paper \cite{20}). Recently Dyakonov and Petrov found much more elegant formula for path ordered exponent. Namely, for the case when $Z$ is Wilson loop and gauge group is $SU(2)$, they derived from (\ref{13}) the following expression for $Z$: \begin{eqnarray} Z&=&\mathop{\mbox{tr P}} e^{ i\oint_{\gamma} \mathop{dx^{\mu}} A_{\mu}(x) }\nonumber\\ &=&\int \mathop{Dn(x)} \prod_{x\in\Sigma} \d(n^a(x) n^a(x) -1) \nonumber\\ &&\exp\left\{ \frac{i}{4}\int_{\Sigma}\mathop{dx^{\mu}\wedge dx^{\nu}} \left[ -F^a_{\mu\nu}n^a +\varepsilon^{abc}n^a D_{\mu}n^b D_{\nu} n^c \right] \right\} \label{14} \end{eqnarray} \noindent where $\Sigma$ is two dimensional surface spanned on contour $\gamma$, $a=1,2,3$, $D^{\mu}$ -- covariant derivative. This formula can be considered as non-Abelian variant of Stokes theorem. We shall continue the discussion of this result in section~4. In the present paper we will derive alternative bosonic path integral representation for $Z$ in the general case $B\in su(N)$. The "action" in this representation is quadratic and so it is much more simpler then one in (\ref{13}) and (\ref{14}). This derivation is presented in the section 2. In the section 3 we check the representation obtained by direct evaluation of the functional integral that defines $Z$. In fact, we give alternative proof of results obtained in the section 2. In the section 4 we derive an non-Abelian analog of Stokes theorem and compare our results with those due to Dyakonov and Petrov. In the section 5 we derive bosonic worldline path integral representation for fermionic determinant and Green functions in Euclidean QCD. In the section 6 we get analogous results for Minkowski space and then, applying stationary phase method, derive quasiclassical equations of motion in QCD. In the last section we summarize our results and discuss perspectives of future investigations. In two appendixes we derive some auxiliary formulae that are used in the main text of the paper. \section{Bosonic path integral representation for the trace of path ordered exponent} Let $N\times N$ matrix $U(t)$, $0\le t\le 1$, be defined by equations \begin{eqnarray} \frac{dU}{dt}&=&iB(t)U(t) \nonumber\\ U(0)&=&I_N \label{15} \end{eqnarray} \noindent where $I_N$ is $N\times N$ unit matrix. Then, obviously, \begin{equation} Z=\mathop{{\rm tr}} U(1) \label{16} \end{equation} First, we consider the case $B \in su(N)$, that is \begin{equation} B^{\dag }=B, \ \ \mathop{{\rm tr}} B=0 \label{16.1} \end{equation} Let us consider an operator \begin{equation} \hat B = a^{\dag}_r B^r_s a^s \label{17} \end{equation} \noindent that acts in some Fock space ${\cal F}$. Operators $a^{\dag}_r$ and $a^s$ in (\ref{17}) are usual {\it bosonic} creation and annihilation ones, $[a^r,a^{\dag}_s]=\delta^r_s$. Let ${\cal H}_n$ be $n$-particle subspace of ${\cal F}$, $\Pi_n$ is orthogonal projector on ${\cal H}_n$. Then \begin{equation} Z= \mathop{{\rm tr}} \Pi_1 \mathop{{\rm P}} e^{i\int_0^1 \mathop{dt} \hat B (t)} \label{18} \end{equation} Indeed, if $\hat N=a^{\dag}_r a^r$, then \begin{equation} \left[ \hat B, \hat N \right] =0 \label{19} \end{equation} So $\hat B {\cal H}_n \subset {\cal H}_n$ and in one dimensional subspace ${\cal H}_1$ the operator $\hat B$ can be identified with the matrix $B$ via relation: \begin{equation} \hat B a^{\dag}_r \phi^r |0> = a^{\dag}_r (B^r_s\phi^s) |0> \label{20} \end{equation} Projector $\Pi_1$ can be represented as \begin{eqnarray} \Pi_1&=&\int_{1-\varepsilon}^{1+\varepsilon \delta(\hat N-\lambda)\nonumber \\ &=&\int_{1-\varepsilon}^{1+\varepsilon\int\limits_{-\infty}^{\infty e^{-i\lambda\eta + i\hat N \eta} \label{21} \end{eqnarray} \noindent where $0<\varepsilon<1$. So \begin{eqnarray} Z&=&\lim_{\delta \to +0} \mathop{{\rm tr}} \Pi_1 e^{-\delta\hat N} \mathop{{\rm P}} e^{i\int_0^1 \mathop{dt} \hat B (t)} \nonumber \\ &=&\lim_{\delta \to +0} \int_{1-\varepsilon}^{1+\varepsilon \int\limits_{-\infty}^{\infty \mathop{{\rm tr}} e^{-i\lambda\eta +(i\eta-\delta)\hat N} e^{i\int_0^1 \mathop{dt} \hat B (t)}\nonumber\\ &=&\lim_{\delta \to +0} \int_{1-\varepsilon}^{1+\varepsilon \int\limits_{-\infty}^{\infty \mathop{{\rm tr}} e^{-i\lambda\eta} \mathop{\mbox{tr P}} e^{i\int_0^1 \mathop{dt} (\hat B (t) +(\eta +i\delta)\hat N)} \label{22} \end{eqnarray} \noindent The latter equality is valid due to (\ref{19}). The trace of ordered exponent in (\ref{22}) can be represented as functional integral. In our case its explicit form essentially depends on some subtleties in its definition. So let me remind in a few words the general construction of functional integral. For more detailed discussion see, for instance, \cite{22}. Let \begin{equation} \check Z=\mathop{\mbox{tr P}} e^{\int^1_0 \mathop{dt} \hat H} \label{23} \end{equation} \noindent where $\hat H=\hat H(a^{\dag},a;t)$ is an operator in Fock space ${\cal F}$. In our case \begin{equation} \hat H=(i\eta-\delta)\hat N +i\hat B \label{24.1} \end{equation} Equivalently, we can write \begin{equation} \hat H=\hat H(\hat p,\hat q;t) \label{24.2} \end{equation} \noindent \begin{equation} \hat p=\frac{a+a^{\dag}}{\sqrt 2}, \qquad \hat q= \frac{a-a^{\dag}}{i\sqrt 2} \label{24} \end{equation} Let $H=H(p,q;t)\equiv H(\bar z,z;t)$ be Weyl, or normal, or any other symbol of the operator $\hat H$, \begin{equation} z=\frac{p-iq}{\sqrt 2},\qquad \bar z=\frac{p+iq}{\sqrt 2} \label{25} \end{equation} Further, let $*$ be the operation that represent the multiplication of operators in the language of symbols. This means that if $\hat K=\hat K(\hat p,\hat q)$ and $\hat L=\hat L(\hat p,\hat q)$ are some operators with symbols $ K=K( p, q)$ and $ L= L( p, q)$ and $\hat M=\hat K \hat L$ then \begin{equation} M(p,q)=(K*L)(p,q) \label{26} \end{equation} In this terms \begin{equation} \check Z=\lim_{n\to \infty}\int\mathop{dpdq}\left( e^{\frac{1}{n}H(\cdot,\cdot;0)}* e^{\frac{1}{n}H(\cdot,\cdot;\frac{1}{n})}*\ldots* e^{\frac{1}{n}H(\cdot,\cdot;\frac{n-1}{n})}\right)(p,q) \label{27} \end{equation} The last formula can be rewritten as functional integral. Its concrete form depends on the choice of a kind of symbols used in (\ref{27}). In particular, for Weyl symbols \begin{equation} \check Z={\cal N}\int_{PBC}DpDq\: e^{\int_0^1 dt\:(ip(t)\dot q(t)+ H_W(p(t),q(t);t))} \label{28} \end{equation} \noindent whereas for normal symbols \begin{equation} \check Z=\lim_{\epsilon \to +0}{\cal N}' \int_{PBC} DzD\bar z \: e^{ \int_0^1 \mathop{dt}(\bar z(t)\dot z(t)+H_{norm}(\bar z(t),z(t+\epsilon))} \label{29} \end{equation} \noindent Here $PBC$ means 'periodic boundary conditions', $z$ and $\bar z$ are {\rm independent} complex variables, and ${\cal N,N'}$ are normalization constants (that we will usually omit in what follows). In our case \begin{eqnarray} H_W(p,q;t)&=&H_W(z^{\dag},z;t)\nonumber\\ &=&(i\eta-\delta)z^{\dag}z+iz^{\dag} Bz-\frac{N}{2}(i\eta-\delta) \label{30} \end{eqnarray} \noindent where $p,q$ are connected with $z$ by formulae (\ref{25}), and \begin{equation} H_{norm}(\hat z,z)=(i\eta-\delta)z^{\dag} z +iz^{\dag} B z \label{31} \end{equation} Formulae (\ref{28}), (\ref{29}) correspond to standard sign conventions. However, for our purposes it is convenient to change variables $$ z(t)\to z(1-t)$$ \noindent Then formulae (\ref{28}), (\ref{29}) with symbols (\ref{30}), (\ref{31}) can be rewritten as \begin{eqnarray} \check Z&=&\int_{PBC} D^2 z\:\exp\left\{ i\int_0^1\mathop{dt} \left[ iz^{\dag} (t) \dot z(t) +z^{\dag} (t) B(t) z(t)\right.\right.\nonumber\\ & +&\left.\left. (\eta +i\delta)z^{\dag}(t)z(t)-\frac{N}{2}(\eta+i\delta)\right]\right\} \label{32} \end{eqnarray} \noindent where $D^2z\equiv D({\rm Re} z) D({\rm Im} z)$ and \begin{eqnarray} \check Z& =&\lim_{\epsilon \to +0} \int DzD\bar z \: \exp\left\{ i\int^1_0 \mathop{dt} \left[ i\bar z(t) \dot z(t) +\bar z(t)B(t)z(t-\epsilon)\right.\right.\nonumber\\ &{}&\left.\left. +(\eta+i\delta)\bar z(t)z(t-\epsilon) \right] \right\} \label{33} \end{eqnarray} \noindent respectively. There are several essential differences between formulae (\ref{32}) and (\ref{33}). First, in (\ref{32}) $z$ and $z^{\dag}$ are complex conjugated variables whereas in (\ref{33}) $\bar z$ and $z$ are independent. The second, the last term in the "action" in (\ref{32}) is absent in (\ref{33}). The third, there is the shift of "time" variable in the last two terms in the "action" in (\ref{33}) that is absent in (\ref{32}). In the next section we will show by explicit calculations that this shift just compensates the absence of the term $$ -\frac{N}{2}(\eta +i\delta)$$ \noindent in (\ref{33}). Finally, it is worth to note that the limit in (\ref{33}) must be evaluated {\it after} functional integration in (\ref{33}) because this two operations don't commute. It will be confirmed by explicit calculations in the section 3. Substituting (\ref{32}) and (\ref{33}) in (\ref{22}) and evaluating limits $\varepsilon \to 0, \ \ \delta \to 0$ (bat not the limit $\epsilon \to +0$!), ones obtains, respectively, \begin{eqnarray} Z&=&\int_{PBC} D^2z \: \int\limits_{-\infty}^{\infty e^{i\int^1_0 \mathop{dt} (iz^{\dag}\dot z+z^{\dag} Bz+ \eta(z^{\dag} z-1-\frac{N}{2})} \label{34} \end{eqnarray} \noindent and \begin{eqnarray} Z&=&\lim_{\epsilon \to +0} \int_{PBC} DzD\bar z \int\limits_{-\infty}^{\infty \exp \left\{ i\int_0^1\mathop{dt} i\bar z \dot z +\bar z B e^{-\epsilon \frac{d}{dt}}z \right. \nonumber\\ &+&\left.\eta(\bar ze^{-\epsilon \frac{d}{dt}}z -1)\right\} \label{35} \end{eqnarray} We won't try to justify here the validity of limiting procedure $\varepsilon \to 0$, $\mbox{$\delta \to 0$}$ because in the next section we will check formulae (\ref{34}), (\ref{35}) by direct calculation. In (\ref{34}) one can integrate with respect to $\eta$: \begin{equation} Z=\Dzi e^{i\int_0^1 \mathop{dt}(iz^{\dag}\dot z +z^{\dag} Bz)} \label{36} \end{equation} But we cannot obtain $\delta$-function by integration with respect to $\eta$ in (\ref{35}) because $\bar z$ and $z$ in (\ref{35}) are independent complex variables. One can also get another useful form of the representation ( \ref{36}), namely \begin{equation} Z=\int_{PBC}D^2z \: \prod \limits _t \delta(z^{\dag} (t)z(t)-1-\frac{N}{2}) e^{i\int_0^1 \mathop{dt}(iz^{\dag}\dot z +z^{\dag} Bz)} \label{37} \end{equation} In what follows, we will sometimes omit symbol $\prod$ in formulae of the type (\ref{37}). To get formula (\ref{37}), it is sufficient to insert projectors $\Pi_1$ represented in the form (\ref{21}) between each pair of adjacent factors in (\ref{27}) and to repeat all calculations that have led us to representation (\ref{36}). The same arguments allow also to obtain an analogous variant of (\ref{35}): \begin{eqnarray} Z&=&\lim_{\epsilon \to +0} \int_{PBC} Dz(t)D\bar z(t)D\eta(t) \exp \left\{ \right. i\int_0^1\mathop{dt} i\bar z (t) \dot z(t) +\bar z(t) B(t) e^{-\epsilon \frac{d}{dt}}z(t) \nonumber\\ &+&\eta(t)(\bar z(t)e^{-\epsilon \frac{d}{dt}}z(t) -1)\left.\right\} \label{37.1} \end{eqnarray} Thus we have derived four different representations (\ref{35})-(\ref{37.1}) for one quantity $Z$. They all appear to be useful in quantum field theory. We have got (\ref{35})-(\ref{37.1}) assuming that $B\in su(N)$. But these representation remains valid for every trace free matrix by virtue of analytical continuation. Representation for $B\in GL(N)$ can be obtained by extracting of the trace part of the matrix $B$ at the beginning of calculations. Indeed, if $$B=B'+ \frac{1}{N}I_N\mathop{{\rm tr}} B$$ \noindent then $\mathop{{\rm tr}} B'=0$ and \begin{equation} \mathop{\mbox{tr P}} e^{i\int_0^1 \mathop{dt} B}= e^{\frac{i}{N}\int_0^1 \mathop{dt} \mathop{{\rm tr}} B} \mathop{\mbox{tr P}} e^{i\int_0^1 \mathop{dt} B'} \label{39} \end{equation} Another representation for matrices with non-zero trace will be given in the next section (see eq. (\ref{58})). Finally, let us derive a kind of representation (\ref{37}) for the case \begin{equation} B(t)=\sum_{j=1}^{M} C_j(t) \otimes D_j(t) \label{39.1} \end{equation} \noindent where $C_j$ and $D_j$ are $N_1\times N_1$ and $N_2\times N_2$ matrices. Let $a_{(i)}^{r_i},\ \ a^{\dag}_{(i)r_i},\ \ i=1,2 \ \ r_i=1,2,\ldots ,N_i$ be two sets of annihilation and creation operators that act in some Fock space, and $\Pi_{1\otimes 1}$ is a projector on the space \begin{equation} {\cal H}_{1\otimes 1}=\left\{\phi^{r_1 r_2}a^{\dag}_{(1)r_1} a^{\dag}_{(2)r_2}|0>\right\} \label{39.2} \end{equation} One can write \begin{eqnarray} \lefteqn{ \mathop{\mbox{tr P}} e^{i\int_0^1 \mathop{dt} B}}\nonumber\\ &&=\mathop{{\rm tr}} \Pi_{1\otimes 1}\: \mbox{P}\: \exp\left\{ i\int^1_0 \mathop{dt} \sum\limits_{j=1}^M(a^{\dag}_{(1)}C_j a_{(1)}) (a^{\dag}_{(2)}D_j a_{(2)})\right\} \label{39.3} \end{eqnarray} \noindent The last formula ia an analog of (\ref{18}). Then, repeating the calculations that have been done for derivation of eq. (\ref{36}) from eq. (\ref{18}), one obtains: \begin{eqnarray} \mathop{\mbox{tr P}} e^{i\int_0^1 \mathop{dt} B} &=&\int D^2z_1 D^2 z_2 \prod\limits_{i=1}^2 \delta (z^{\dag}_{(i)}z_{(i)}-1-\frac{N}{2}) \nonumber\\ &&\exp \left\{ i\int_0^1 \mathop{dt} \left[ i\sum\limits_{i=1}^2 z^{\dag}_{(i)}\dot z_{(i)}\right.\right.\nonumber\\ &&\left.\left. +\sum\limits_{j=1}^M (z^{\dag}_{(1)}C_j z_{(1)} -\frac{1}{2}\mathop{{\rm tr}} C_j) (z^{\dag}_{(2)}D_j z_{(2)}-\frac{1}{2}\mathop{{\rm tr}} D_j) \right] \right\} \label{39.4} \end{eqnarray} Similar arguments allow to obtain the analog of eqs. (\ref{35}), (\ref{37}), and (\ref{37.1}) for an matrix $B$ defined by eq. (\ref{39.1}). \section{Alternative proof of bosonic path integral representation for the trace of path ordered exponent} At first, we will evaluate the integral (\ref{36}) assuming that $B\in su(N)$ . One can write: \begin{eqnarray} Z&=&\lim_{\epsilon \to +0} \Dzi e^{-\epsilon\int^1_0 \mathop{dt} z^{\dag} z}e^{i\int^1_0 \mathop{dt}(iz^{\dag} \dot z +z^{\dag} B z)}\nonumber\\ &=&\lim_{\epsilon \to +0} \int\limits_{-\infty}^{\infty e^{-i\eta\left( 1+\frac{N}{2}\right)}\int_{PBC} D^2z \: e^{\int^1_0 \mathop{dt} z^{\dag}\left(-\frac{d}{dt} +iB+i\eta -\epsilon\right)z} \label{40} \end{eqnarray} Performing functional integration, one gets: \begin{equation} Z=\lim_{\epsilon \to +0} \int\limits_{-\infty}^{\infty \frac{e^{-i\eta\left( 1+\frac{N}{2}\right)}}{ {\det}\left(-\frac{d}{dt} +i(B+\eta)-\epsilon\right)_{PBC}} \label{41} \end{equation} To evaluate integral in (\ref{41}), let us consider eigenvalue problem \begin{equation} \left(-\frac{d}{dt} +i(B+\eta)-\epsilon\right)\phi (t)=\lambda \phi(t) \label{42} \end{equation} \begin{equation} \phi(0)=\phi(1) \label{43} \end{equation} The general solution of eq. (\ref{42}) is \begin{equation} \phi=e^{(-\lambda +i\eta -\epsilon)t} U(t)\chi \label{44} \end{equation} \noindent where $U(t)$ is a solution of (\ref{15}) and vector $\chi$ doesn't depend on $t$. Let $e^{i\alpha_r}$ and $\xi_r,\ \ r=1,\ldots,N$ be eigenvalues and eigenvectors of the matrix $U(1)$: \begin{equation} U(1)\xi_r=e^{i\alpha_r}\xi_r \label{45} \end{equation} One notes that $U(1)\in SU(N)$ and so $\alpha_r$ are real and \begin{equation} \sum_{r=1}^N \alpha_r=0 \label{46} \end{equation} Further, to satisfy (\ref{43}) we must put \begin{equation} \lambda\equiv \lambda_{rn}=-\epsilon +i\eta +i\alpha_r +2\pi i n,\ \ n=0,\pm 1,\ldots \label{47} \end{equation} \noindent and $\chi=\xi_r$. So \begin{eqnarray} \lefteqn{{\det}\left(-\frac{d}{dt} +i(B+\eta)-\epsilon\right)_{PBC}= \prod\limits_{r=1}^{N} \prod\limits_{n=-\infty}^{\infty}(2\pi i n+i\alpha_r +i\eta-\epsilon)} \nonumber\\ &=&\left[-i\prod_{n=-\infty}^{\infty}(2\pi in)\right]^r \prod\limits_{r=1}^{N} (\alpha_r+\eta +i\epsilon)\prod\limits_{n\ne 0} \left(1+\frac{\alpha_r +\eta +i\epsilon}{2\pi n}\right)\nonumber\\ &=&\left[-i\prod_{n=-\infty}^{\infty}(2\pi in)\right]^r \prod\limits_{r=1}^{N}(\alpha_r+\eta +i\epsilon) \prod\limits_{n=1}^{\infty} \left( 1-\frac{(\alpha_r+\eta +i\epsilon)^2}{4\pi^2n^2}\right) \label{48} \end{eqnarray} \noindent Then, omitting irrelevant infinite constant and using well-known formula \begin{equation} \prod\limits_{n=1}^{\infty}\left( 1-\frac{a^2}{n^2}\right)=\frac{1}{\pi a} \sin \pi a \label{49} \end{equation} \noindent one gets \begin{equation} {\det}\left(-\frac{d}{dt} +i(B+\eta)-\epsilon\right)_{PBC}= \prod_{r=1}^{N}\sin \left( \frac{\alpha_r+\eta+i\epsilon}{2}\right) \label{50} \end{equation} Substituting (\ref{50}) in (\ref{41}), one obtains contour integral \begin{equation} Z=\lim_{\epsilon \to +0} \int\limits_{-\infty}^{\infty \frac{e^{-i\eta\left(1+\frac{N}{2}\right)}}{ \prod_{r=1}^{N}\sin \left( \frac{\alpha_r+\eta+i\epsilon}{2}\right) } \label{51} \end{equation} One can close the contour of integration in (\ref{51}) in the lower half plane and represent $Z$ as the sum of residues in the poles \begin{equation} \eta_{rn}=-\alpha_r -i\epsilon +2\pi n \label{52} \end{equation} The contribution of the pole with given $r$ and $n$ is \begin{equation} 2\pi i(-1)^{n+1}\frac{e^{i(\alpha_r+i\epsilon)}}{ \prod\limits_{{1\le s\le N \atop s\ne r}}\sin \left( \frac{\alpha_s-\alpha_r}{2}\right) } \label{53} \end{equation} \noindent So, omitting again inessential constant, one gets: \begin{eqnarray} Z&=&\lim_{\epsilon \to +0} \sum\limits_{r,n} \left( \begin{array}{c} \mbox{contribution of the residue in}\\ \eta_{rn}=-\alpha_r -i\epsilon +2\pi n \end{array} \right) \nonumber\\ &=& \sum\limits_{r=1}^N\frac{e^{i\alpha_r(1+\frac{N}{2})}}{ \prod\limits_{ {\scriptstyle {s=1 \atop s\ne r}} }^N \sin \left( \frac{\alpha_s-\alpha_r}{2} \right)} \label{54} \end{eqnarray} \noindent Finally, using elementary but rather non-obvious identity \begin{equation} \sum\limits_{r=1}^N\frac{e^{i\alpha_r(1+\frac{N}{2})}}{ \prod\limits_{ {\scriptstyle {s=1\atop s\ne r}}} ^N \sin \left( \frac{\alpha_s-\alpha_r}{2} \right)} = (-2i)^{N-1} \left(e^{\frac i 2 \sum\limits_{r=1}^N \alpha_r}\right) \sum\limits_{r=1}^N e^{i\alpha_r} \label{55} \end{equation} \noindent (that is valid for any complex numbers $\alpha_r$; see Appendix A for proof), one obtains: \begin{equation} Z= \sum\limits_{r=1}^N e^{i\alpha_r}=\mathop{{\rm tr}} U(1) =\mathop{\mbox{tr P}} e^{i \int^1_0 \mathop{dt} B(t)} \label{56} \end{equation} \noindent This proves the representation (\ref{36}). Throughout the proof we didn't control inessential numerical normalization factors arising in front of functional integrals, determinants, etc. They can be easily reconstructed from normalization condition \begin{equation} Z\raisebox{-10pt}{\rule{0.4pt}{20pt}${\scriptstyle B=0}$} =N \label{57} \end{equation} The representation (\ref{36}) is valid for traceless matrix $B$. In general case the following representation is valid \begin{equation} Z=e^{-\frac i 2 \int^1_0 \mathop{dt} \mathop{{\rm tr}} B(t)} \int D^2z \: \delta \left( \int^1_0 \mathop{dt} z^{\dag} z -1- \frac N 2\right) e^{i\int_0^1 \mathop{dt}(iz^{\dag}\dot z +z^{\dag} Bz)} \label{58} \end{equation} \noindent This representation is alternative to one given by eq. (\ref{39}). To prove (\ref{58}), it is sufficient to repeat the proof given for representation (\ref{36}) but without using the condition (\ref{46}) . One can trace the cancellation of the pre-integral factor in (\ref{58}) and contribution of the factor $$ e^{\frac i 2 \sum\limits_{r=1}^N \alpha_r} $$ \noindent in (\ref{55}). Now let us check the representation (\ref{37}). Representing $\delta$-function in (\ref{37}) as \begin{equation} \prod_t \delta \left( z^{\dag} z -1 -\frac N 2 \right)=\int D\eta \: e^{i\int^1_0 \mathop{dt} \left( z^{\dag} z -1 -\frac N 2 \right)\eta } \label{60} \end{equation} \noindent and integrating over $z^{\dag}, z$, one obtains the analog of (\ref{41}): \begin{equation} Z=\lim_{\epsilon \to +0} \int D\eta(t)\: \frac{e^{-i\left( 1+\frac{N}{2}\right) \int^1_0 \mathop{dt} \eta(t)}}{ {\det}\left(-\frac{d}{dt} +i(B(t)+\eta(t))-\epsilon\right)_{PBC}} \label{61} \end{equation} The eigenvalue problem \begin{equation} \left(-\frac{d}{dt} +i(B(t)+\eta(t))-\epsilon\right)\phi (t)=\lambda \phi(t) \label{62} \end{equation} \begin{equation} \phi(0)=\phi(1) \label{62.1} \end{equation} \noindent has the solution \begin{equation} \lambda\equiv \lambda_{rn}=-\epsilon +i\alpha_r +i\int^1_0 \mathop{dt} \eta(t) +2\pi i n,\ \ n=0,\pm 1,\ldots \label{63} \end{equation} \noindent where numbers $\alpha_r$ are defined by equation (\ref{45}). So determinant in (\ref{61}) depends on $\eta(t)$ only via \begin{equation} \eta=\int^1_0 \mathop{dt} \eta(t) \label{64} \end{equation} \noindent Therefore, if one introduce new variables $\eta$ and \begin{equation} \eta_n=\int^1_0 \mathop{dt} e^{2\pi int} \eta(t),\ \ n\ne 0 \label{65} \end{equation} \noindent instead of $\eta(t)$ in (\ref{61}), one finds that integration with respect to $\eta_n$ gives only inessential constant and so the integral (\ref{61}) transforms just in the integral (\ref{41}). But this means that we reduced the proof of representation (\ref{37}) to one of the representation (\ref{36}) which we have already proved. Finally, let us prove the validity of the representations (\ref{35}) and (\ref{37.1}). The proof can be reduced again to one of the representation (\ref{36}) by means of the formula \begin{equation} \lim_{\epsilon \to +0} {\det} \oee =e^{\frac i 2 \int^1_0 \mathop{dt} \mathop{{\rm tr}} B(t)} {\det} \left( -\frac{d}{dt} +iB(t) \right) \label{66} \end{equation} \noindent that is valid for any $N\times N$ matrix $B(t)$ (the proof of (\ref{66}) is given in Appendix~B). Indeed, evaluating the integral with respect to $\bar z, \, z$ in (\ref{35}), one obtains, after inserting of the factor $\exp\left( -\delta \int^1_0 \mathop{dt} z^{\dag} (t)z(t+\epsilon)\right)$, \begin{equation} Z=\lim_{\delta \to +0}\lim_{\epsilon \to +0} \int\limits_{-\infty}^{\infty \frac{e^{-i\eta }}{ {\det}\left(-\frac{d}{dt} +i(B+\eta+i\delta)e^{-\epsilon\frac{d}{dt}}\right)_{PBC}} \label{67} \end{equation} \noindent and further application of (\ref{66}) reduce (\ref{67}) to (\ref{41}). (Remind, that $\mathop{{\rm tr}} B=0$ by assumption.) The same arguments allow to check the representation(\ref{37.1}). Thus we have checked all four representations for the trace of path ordered exponent derived in the section 2. We see the proofs given in the present section are very unlike ones given in the section 2. This can be considered as strong confirmation of the validity of the representations obtained. \section{A variant of non-Abelian Stokes theorem} Analogs of Stokes theorem for Wilson loop \begin{equation} Z=\wl \label{68} \end{equation} \noindent were proposed by several authors \cite{23}. In this works Wilson loop is expressed via some area integral over surface spanned on $\gamma$ with rather complicated path ordered prescriptions. Recently Dyakonov and Petrov derived another, very smart variant of non-Abelian Stokes theorem. Their result we already cited (see (\ref{14})). In this section we will prove a new variant of non-Abelian Stokes theorem for $A_{\mu}\in su(N)$ that is similar to one given by Dyakonov and Petrov. For the case $N=2$ we will be able to derive from our results the Dyakonov-Petrov's formula (\ref{14}) but in slightly corrected form. The little discrepancy between our results and those by Dyakonov and Petrov arises, most likely, because of some subtleties in the definition of the corresponding functional integrals. Let the path $\gamma$ in (\ref{68}) be parametrized as \begin{equation} \gamma=\{ x^{\mu}=q^{\mu}(t),\:0\le t\le 1,\: q^{\mu}(0)=q^{\mu}(t)\} \label{69} \end{equation} \noindent Then \begin{equation} Z=\mathop{\mbox{tr P}} e^{i\int^1_0 \mathop{dt} \dot q ^{\mu}(t) A_{\mu}(q(t))} \label{70} \end{equation} \noindent and we can apply the representation (\ref{37}): \begin{equation} Z= \int D^2\psi \delta (z^{\dag} z -1- \frac N 2) e^{\int^1_0 \mathop{dt} z^{\dag}\left(-\frac{d}{dt} +i\dot q A\right)z} \label{71} \end{equation} Let $\xi^r=\xi^r(x), \: r=1,\ldots,N$ be any field such that \begin{equation} \xi^r(q(t))=z^r(t) \label{72} \end{equation} \noindent The formula (\ref{71}) can be rewritten as \begin{equation} Z= \int \mathop{D^2\xi} \prod\limits_{x\in\gamma}\delta\left( \xi^{\dag}(x) \xi(x)-1-\frac N 2 \right) e^{-\oint_{\gamma} dx^{\mu} \: \xi^{\dag}D_{\mu}\xi(x)} \label{73} \end{equation} \noindent where $D_{\mu}=\partial_{\mu}-iA_{\mu}$ is the usual covariant derivative. Further, applying the classical Stokes theorem, one can easy to prove that \begin{equation} \oint_{\gamma} dx^{\mu} \: \xi^{\dag}D_{\mu}\xi(x) = \int_{\Sigma} \mathop{dx^{\mu}\wedge dx^{\nu}} \left(D_{\mu}\xi^{\dag}D_{\nu}\xi -\frac i 2 \xi^{\dag}F_{\mu\nu}\xi \right) \label{74} \end{equation} \noindent where $\Sigma$ is any surface for which $\partial \Sigma=\gamma$. Substituting (\ref{74}) in (\ref{73}), we get our variant of non-Abelian Stokes theorem: {\samepage \begin{eqnarray} Z \equiv&&\wl\nonumber\\ =&& \int \mathop{D^2\xi} \prod\limits_{x\in\gamma}\delta\left( \xi^{\dag}(x) \xi(x)-1-\frac N 2 \right) e^{ \int_{\Sigma} \mathop{dx^{\mu}\wedge dx^{\nu}} \left(-D_{\mu}\xi^{\dag}D_{\nu}\xi +\frac i 2 \xi^{\dag}F_{\mu\nu}\xi \right) }\nonumber\\ \label{75} \end{eqnarray} } Now let us transform eq. (\ref{75}) into the form that would be similar to eq. (\ref{14}). Let $\lambda^a,\: a=1,\ldots,N^2-1$ be Hermitian generators of $SU(N)$ in the fundamental representation. They can be normed as \begin{equation} \mathop{{\rm tr}} \lambda^a\lambda^b=2\delta^{ab} \label{76} \end{equation} \noindent and satisfy equations \begin{equation} [\lambda^a, \lambda^b] =2 i f^{abc}\lambda^c \label{77} \end{equation} \begin{equation} [\lambda^a, \lambda^b]_{+} = \frac 4 N \delta^{ab}I_N +2d^{abc}\lambda^c \label{78} \end{equation} One notes that \begin{equation} A_{\mu}=\frac{\lambda^a}{2}A_{\mu}^a,\: \ F_{\mu\nu}= \frac{\lambda^a}{2}F_{\mu\nu}^a= \frac{\lambda^a}{2}(\partial_{\mu}A^a_{\nu}- \partial_{\nu}A^a_{\mu} +f^{abc}A^b_{\mu}A^c_{\nu}) \label{78'} \end{equation} Matrices $\lambda^a$ also obey "Fierz" identities: \begin{eqnarray} \frac 1 2 \lambda^a_{ij}\lambda^a_{kl}+ \frac 1 N \delta_{ij}\delta_{kl}&=& \delta_{il}\delta_{kj} \label{79}\\ if^{abc}\lambda^a_{ij}\lambda^b_{kl}\lambda^c_{mn}&=& 2(\delta_{il}\delta_{mj}\delta_{km}- \delta_{in}\delta_{jk}\delta_{ml}) \label{80} \end{eqnarray} One introduces new variables \begin{equation} I^a(x)=\xi^{\dag}(x)\lambda^a\xi(x) \label{81} \end{equation} Variables $I^a$ are not independent. Using "Fierz" identity (\ref{79}), formula (\ref{81}) can be represented as \begin{eqnarray} \frac{1}{\xi^{\dag} \xi}(I^a\lambda^a)_{ij}+\frac 1 N \delta_{ij}= \xi_i\xi^{\dag}_j\frac{1}{(\xi^{\dag}\xi)} \label{82} \end{eqnarray} Let $I\equiv I^a\lambda^a, \: \xi^{\dag}\xi\equiv c$. One notes that $c=1+\frac N 2$ on the surface of integration in (\ref{75}). The matrix $I$ can be represented in the form \begin{equation} I=U diag(a_1,\ldots,a_N)U^{\dag} \label{83} \end{equation} \noindent where \begin{equation} \sum_{i=1}^{N}a_i=0,\ \ U\in SU(N) \label{84} \end{equation} The matrix in R.H.S. of eq. (\ref{82}) is a projector on the one dimensional subspace. This means that eigenvalues of the matrix in the L.H.S. of eq. (\ref{82}) are all equal to zero except only one that is equal to 1. So, up to renumbering of eigenvalues, \begin{eqnarray} a_i&=&-\frac{2c}{N},\ \ i=1,\ldots,N-1\nonumber\\ a_N&=&2c\frac{(N-1)}{N} \label{85} \end{eqnarray} One notes that \begin{eqnarray} \mathop{{\rm tr}} I^k&=&(N-1)\left( -\frac{2c}{N}\right)^k +\left[ 2c\frac{(N-1)}{N}\right]^k\nonumber\\ &=&\frac{(2c)^k (N-1)}{N^k} \left[ (N-1)^{k-1} +(-1)^k \right] \label{94} \end{eqnarray} Thus the matrix $I$ has $N-1$ coincident eigenvalues. But this means that the solutions of eq. (\ref{82}) are in one-to-one correspondence with the points of a coset $SU(N)/U(1)\times SU(N-1)${} $\approx CP^{N-1}$. Indeed, the matrix $U$ in formula (\ref{83}) can be presented in the form \begin{equation} U=U_1V_1V_2 \label{85.1} \end{equation} \noindent where \begin{eqnarray} V_1= \left( \begin{array}{cc} \tilde{V}_1& \begin{array}{c} 0\\ \vdots\\ 0 \end{array}\\ 0\ldots 0&1 \end{array} \right), \ \ \tilde V_1 \in SU(N-1) \label{85.2} \\ V_2=\exp \{ diag(t,\ldots,t,-(N-1)t \}\nonumber \end{eqnarray} \noindent and a matrix $U_1$ represents an element of the coset $SU(N)/U(1)\times SU(N-1)${} . So, due to coincidence of the first $N-1$ eigenvalues of the matrix $I$, one finds \begin{equation} I=U_1 diag\left(-\frac{2c}{N},\ldots,-\frac{2c}{N}, 2c\frac{(N-1)}{N}\right) U^{\dag}_1 \label{86} \end{equation} \noindent Therefore the set of solutions of eq. (\ref{82}) is isomorfic to coset $SU(N)/U(1)\times SU(N-1)${} . Further, using identities (\ref{79}), (\ref{80}), one finds that on the surface $\xi^{\dag} \xi$$=c=const$ \begin{eqnarray} D_{\mu}\xi^{\dag} D_{\nu}\xi&=&\frac{1}{8c^2}\mathop{{\rm tr}} ID_{[\mu}ID_{\nu]}I \label{87}\\ \xi^{\dag} F_{\mu\nu}\xi&=&\frac 1 2 \mathop{{\rm tr}} IF_{\mu\nu} \label{88} \end{eqnarray} \noindent where \begin{equation} D_{\mu}I=\partial_{\mu}I-i[A_{\mu},I] \label{89} \end{equation} Now let us insert in (\ref{75}) an identity \begin{equation} 1=\prod_{x\in\Sigma}\int\prod_a dI^a(x)\delta(\xi^{\dag}(x) \lambda^a \xi(x) -I^a(x)) \label{90} \end{equation} \noindent By virtue of (\ref{87}), (\ref{88}), the eq. (\ref{75}) can be rewritten in the form \begin{equation} Z=\int D\mu(I)\exp\left\{ -i\int\limits_{\Sigma}\mathop{dx^{\mu}\wedge dx^{\nu}}\left[ \frac{1}{8c^2}\mathop{{\rm tr}} ID_{[\mu}ID_{\nu]}I-\frac 1 4 \mathop{{\rm tr}} IF_{\mu\nu}\right]\right\} \label{91} \end{equation} \noindent where \begin{equation} D\mu(I)=\left(\prod_a DI^a\right) \int D^2\xi \delta(\xi^{\dag} \xi- 1-\frac N 2)\prod_a\delta(\xi^{\dag} \lambda^a \xi-I^a) \label{92} \end{equation} The consideration given above shows that $D\mu(I)$ is nothing but $SU(N)$-invariant measure on the coset $SU(N)/U(1)\times SU(N-1)${}{} The explicit form of the measure $D\mu(I)$ is rather cumbersome but, fortunately, it appears that measure $D\mu(I)$ in (\ref{91}) can be replaced by the measure \begin{equation} D\mu'(I)\equiv \left(DI^a\right) \prod_{k=2}^N \delta(\mathop{{\rm tr}} I^k- c_{k,N}) \label{93} \end{equation} \noindent where numbers $c_{k,N}$ are defined by formula (\ref{94}) in which one must put $2c=N+2$ by virtue of the first $\delta$-function in eq.(\ref{92}). Indeed, numbers $\mathop{{\rm tr}} I^k$ define uniquely characteristic equation for the matrix $I$ and, consequently, its eigenvalues. This leads to representation (\ref{86}). So on the surface $\mathop{{\rm tr}} I^k=c_{k,N}$ any functional $\Phi(I)$ is invariant under transformation \begin{equation} \Phi(I)\to \Phi(V_1 V_2 IV^{\dag}_2V^{\dag}_1) \end{equation} \noindent where a matrix $V_1 V_2$ is defined by eqs. (\ref{85.2}). Therefore in any integral \begin{equation} \int D\mu'(I)\Phi(I) \end{equation} \noindent one can perform all integrations except those corresponding to integration over coset $SU(N)/U(1)\times SU(N-1)${} : \begin{equation} \int D\mu'(I)\Phi(I) =const \int D\mu(I)\Phi (U_1diag(-\frac{2c}{N},\ldots,-\frac{2c}{N}, 2c\frac{N-1}{N})U^{\dag}_1) \end{equation} \noindent (see (\ref{86})). But this just means, in particular, that one can replace the measure $D\mu(I)$ in (\ref{91}) by the measure $D\mu'(I)$ defined by eq. (\ref{93}). Now we can formulate our final result: \begin{eqnarray} \lefteqn{ \mathop{\mbox{tr P}} \exp\left\{i\oint_{\gamma} \mathop{dx^{\mu}} A_{\mu} \right\}}\nonumber\\ &=&\int DI\prod_{k=2}^N \delta(\mathop{{\rm tr}} I^k-c_{k,N})\nonumber\\ &&\exp \left\{ i\int_{\Sigma}\mathop{dx^{\mu}\wedge dx^{\nu}}\left[ -\frac{1}{2(N+2)^2}\mathop{{\rm tr}} (ID_{\mu}ID_{\nu}I) +\frac 1 4 \mathop{{\rm tr}} IF_{\mu\nu} \right]\right\}\nonumber\\ &&\label{94'} \end{eqnarray} \noindent where \begin{equation} c_{k,N}=\frac{(N-1)(N+2)^k}{N^k}\left[ (N-1)^{k-1} +(-1)^k \right] \label{95} \end{equation} Let us compare our results with those due to Dyakonov and Petrov \cite{21}. Putting in (\ref{94'}) $N=2,\ \ I=I^a\sigma^a$ and the changing $I^a\to 2I^a$ , we get: \begin{eqnarray} \lefteqn{ \mathop{\mbox{tr P}} \exp\left\{i\oint_{\gamma} \mathop{dx^{\mu}} A_{\mu} \right\}}\nonumber\\ &=& \int DI \delta(I^2-1)\exp\left\{-\frac i 2 \int \mathop{dx^{\mu}\wedge dx^{\nu}} \left( \varepsilon^{abc}I^aD_{\mu}I^b D_{\nu}I^c-I^a F^a_{\mu\nu} \right)\right\}\nonumber\\ &&\label{97} \end{eqnarray} Comparing formulae (\ref{97}) and (\ref{14}), we see that they differ by the factor 1/2 in front of the "action". Most likely this discrepancy arises because of some subtleties in the definition of the functional integrals that play the important role in our discussion. In particular, if one ignores the difference between functional integrals constructed by means of Weyl and normal symbols, one obtains the following representation instead of (\ref{73}): \begin{equation} Z= \int \mathop{D^2\xi} \prod\limits_{x\in\gamma}\delta\left( \xi^{\dag}(x) \xi(x)-1 \right) e^{-\oint_{\gamma} dx^{\mu} \: \xi^{\dag}D_{\mu}\xi(x)} \label{98} \end{equation} \noindent Further, using formulae (\ref{87}) and (\ref{94}) with $c=1$ instead of $c=1+\frac N 2$, one can easy trace that representation (\ref{98}) leads exactly to Dyakonov-Petrov formula (\ref{14}). But representation (\ref{98}) is wrong. So just formula (\ref{97}) must be considered as correct version of non-Abelian Stokes theorem for $N=2$. This doesn't mean, however, that Dyakonov-Petrov formula (\ref{14}) is incorrect. But it means that the definition of functional integral in (\ref{14}) must be clarified. Another discrepancy between our results and those due to Dyakonov and Petrov arises in the case $N\ge 3$. Dyakonov and Petrov pointed out in their work \cite {21} that integration in the functional integral representation for Wilson loop must be performed over the coset $SU(N)/[U(N)]^{N-1}$ whereas in our representation (\ref{94'}) integration is carried out, in fact, over the coset $SU(N)/U(1)\times SU(N-1)${} . But nowadays it is hard to discuss this discrepancy because no explicit formulae for the case $N\ge 3$ were given in \cite{21}. {\hsize=13.26cm {\section{Bosonic worldline path integral representation for fermionic determinants and Green functions in Euclidean space}} Bosonic worldline path integral representation for fermionic determinants can be obtained directly from formulae (\ref{8}), (\ref{10}) and results of the section 2: \begin{eqnarray} \lefteqn{ \ln{\det}(i\hat{\nabla}+im)}\nonumber\\ &=&\dT\dQ\nonumber\\ && {\cal N}_T\exp\left\{ \int^1_0 \mathop{dt} \left[ -\frac{\dot q^2}{4T}-z^{\dag} \dot z -\psi^{\dag}\dot{\psi}+i\dot q z^{\dag} A_{\mu} z\right.\right.\nonumber\\ &&\left.\left. -T(\psi^{\dag}\sigma^{\mu\nu}\psi)(z^{\dag} F_{\mu\nu} z) \vphantom{ \int^1_0 \mathop{dt} \left[ -\frac{\dot q^2}{4T}-z^{\dag} \dot z\right. } \right] \right\} \label{99} \end{eqnarray} Here variables $z=\{z^r, \ \ r=1,\ldots,N\}$ and $\psi=\{\psi^i,\: i=1,,2,3,4\}$ describe colour and spin degrees of freedom respectively, $z^{\dag}$ and $\psi^{\dag}$ are complex conjugated to $z$ and $\psi$, \begin{eqnarray} D^2 z&\equiv&\prod\limits_t \prod\limits_{r=1}^N \mathop{d(\mbox{Re}\, z^r(t))}\mathop{d(\mbox{Im}\, z^r(t))}\label{100}\\ D^2 \psi&\equiv&\prod\limits_t \prod\limits_{r=1}^N \mathop{d(\mbox{Re}\, \psi^i(t))}\mathop{d(\mbox{Im}\, \psi^i(t))} \label{100.1} \end{eqnarray} \noindent and ${\cal N}_T$ is a normalization constant. The latter can be evaluated from the condition \begin{equation} <x|\mathop{{\rm tr}} e^{-T\wH}|x>\raisebox{-10pt}{\rule{0.4pt}{20pt}${\scriptstyle A=0}$}= 4N<x|e^{-T\partial_{\mu}\partial_{\mu}}|x>=\frac{N}{2\pi^2T^2} \label{101} \end{equation} Indeed, putting $A=0$ in (\ref{99}) and comparing the result with (\ref{101}), one obtains: \begin{eqnarray} {\cal N}_T^{-1} &=&\left( \frac{N}{2\pi^2T^2}\right)^{-1} \int\limits_{q(0)=q(1)=x} Dq\:\int_{PBC}D^2\psi D^2 z\delta\left( z^{\dag}z -1-\frac N 2 \right)\nonumber\\ &&\delta\left( \psi^{\dag}\psi -3\right) \exp\left\{ \int^1_0 \mathop{dt} \left[ -\frac{\dot q^2}{4T}-z^{\dag} \dot z -\psi^{\dag}\dot{\psi}\right]\right\} \label{102} \end{eqnarray} \noindent Obviously, ${\cal N}_T$ doesn't depend on $x$. Remind, that in Euclidean space $\psi^{\dag}\psi$ is $SO(4)$ scalar. So representation (\ref{99}) is manifestly relativistic and gauge invariant. Our next task is to derive bosonic worldline path integral representation for Euclidean Green functions. In what follows, we restrict ourselves only to derivation of such representation for generating functional $Z(j)$ for vacuum correlators \begin{equation} <\psi^{\dag}_{f_1}(x_1)\psi_{f_1}(x_1)\ldots \psi^{\dag}_{f_n}(x_n)\psi_{f_n}(x_n)> \label{103} \end{equation} \noindent However, our method is quite general and can be easily applied to derivation of analogues representations for arbitrary Green functions. Standard functional integral representation for $Z(j)$ can be written as \begin{equation} Z(j)=\int DA\: e^{S_{YM}}\int D\Psi D\bar{\Psi} e^{\left\{ \sum_f\int dx \:\left[ i\bar{\Psi}_f \hat{\nabla}\Psi_f+im_f \bar{\Psi}_f\Psi_f-ij_f\bar{\Psi}_f\Psi_f\right]\right\}} \label{104} \end{equation} Integrating with respect to fermionic fields, we get \begin{equation} Z(j)=\int DA\: e^{S_{YM}}\prod_f{\det}(i\hat{\nabla}+im_f-ij_f) \label{105} \end{equation} So our task is reduced to derivation of bosonic path integral representation for determinant $$ {\det}(i\hat{\nabla}+im-ij) $$ The latter problem can be easily solved by applying of the results obtained in the section 2. Indeed, \begin{eqnarray} \lefteqn{ \ln{\det} (i\hat{\nabla}+im-ij)=\frac 1 2 \ln[{\det}(i\hat{\nabla}+im-ij)\gamma^5]^{2}}\nonumber\\ &=&\frac 1 2 \ln{\det}(\nabla_{\mu}\nabla^{\mu}-\sigma^{\mu\nu}F_{\mu\nu} +\hat{\partial}j+(m-j)^2)\nonumber\\ &=&\dT\mathop{{\rm tr}} e^{-T (-\nabla_{\mu}\nabla^{\mu}+\sigma^{\mu\nu}F_{\mu\nu}-\hat{\partial} j+2mj-j^2)} \label{105'} \end{eqnarray} Using representation for path ordered exponent from section 2, one obtains: \begin{eqnarray} \lefteqn{\ln{\det}(i\hat{\nabla}+i(m+j))}\nonumber\\ &=& \dT\dQ \nonumber\\ &&\exp\left\{ \int^1_0 \mathop{dt} \left[ -\frac{\dot q^2}{4T}-z^{\dag} \dot z -\psi^{\dag}\dot{\psi}+i\dot q z^{\dag} A_{\mu} z\right.\right.\nonumber\\ &&\left.\left.-T(\psi^{\dag}\sigma^{\mu\nu}\psi)(z^{\dag} F_{\mu\nu} z -\psi^{\dag}\gamma^{\mu}\psi\partial_{\mu}j(q)-2mj(q)+j^2(q) \right] \right\} \label{106} \end{eqnarray} Substituting (\ref{105'}) in (\ref{106}) for each $f$, we get worldline bosonic path integral representation for generating functional $Z(j)$. The corresponding formulae for $n$-point correlators are rather cumbersome but quite computable. Author hopes that they can be used for computer simulations on the lattice. {\hsize=13.26cm \section{Bosonic worldline path integral representation for fermionic determinants and Green functions in Minkowski space and quasiclassical approximation in QCD} } The derivation of bosonic worldline path integral representation for fermionic determinants in Minkowski space is slightly more involved then one in Euclidean space. The origin of complications is non-unitarity of finite dimensional representation of the Lorentz group. The analog of the representation (\ref{8}) in Minkowski space can be written as \begin{equation} \ln{\det}(i\hat{\nabla}-m)=\dTM \mathop{{\rm tr}} e^{iT\wH} \label{107} \end{equation} Trace in (\ref{107}) can be represented as functional integral: \begin{equation} \mathop{{\rm tr}} e^{iT\wH}=\int_{PBC} Dq \: \mathop{\mbox{tr P}} e^{i\int^1_0 \mathop{dt} \left( -\frac{\dot q^2}{4T}+\dot q A(q)+T \sigma^{\mu\nu}F_{\mu\nu}(q)\right)} \label{108'} \end{equation} \noindent eq. (\ref{108'}) is an analog of eq. (\ref{10}). Path ordering in (\ref{108'}) corresponds to colour and spinor structures. However, in contrast to Euclidean case, we cannot directly use the representations of the type (\ref{37}), (\ref{39.4}) to write ordered exponent in (\ref{108'}) as functional integral. Indeed, those representation comprise, in particular, the factor \begin{equation} \delta(\psi^{\dag}\psi-3) \label{108''} \end{equation} \noindent (see (\ref{99})) that is not Lorentz invariant because spinor representations of Lorentz group are not unitary. To obtain manifestly Lorentz invariant representation, we will use, at first, representation (\ref{37}) for describing of colour degrees of freedom and representation (\ref{37.1}) for describing of spinor ones. In such terms eq. (\ref{108'}) can be rewritten as \begin{eqnarray} \lefteqn{\mathop{{\rm tr}} e^{i\wH}}\nonumber\\ &=&\lim_{\epsilon \to +0} \int_{PBC}DqD^2zD\bar{\psi}D\psi D\lambda \delta(z^{\dag} z-1-\frac N 2) \nonumber\\ &&{\cal N}_T \exp\left\{i\int^1_0\mathop{dt}\left[-\frac{\dot q^2}{4T}+iz^{\dag} z +i\bar{\psi}\dot{\psi}+\right.\right.\nonumber\\ &&\left.\left.z^{\dag} \dot q A(q) z+ T(\bar{\psi}\sigma^{\mu\nu}e^{-\epsilon \frac d{dt}}\psi)(z^{\dag} F^{\mu\nu}z)+\lambda(\bar{\psi}e^{-\epsilon\frac{d}{dt}} \psi-1)\right]\right\} \label{108} \end{eqnarray} In eq. (\ref{108}) the measure $D^2z$ is defined by (\ref{100}) but $\psi$ and $\bar{\psi}$ are independent complex variables. ${\cal N}_T$ is a normalization constant that will be computed later. Integrating over $\bar{\psi},\ \ \psi$ in (\ref{108}), one gets: \begin{eqnarray} \lefteqn{\mathop{{\rm tr}} e^{i\wH}}\nonumber\\ &=&\lim_{\epsilon \to +0} \int_{PBC}DqD^2z D\lambda \delta(z^{\dag} z-1-\frac N 2) \nonumber\\ &&{\cal N}_T {\det}^{-1}\left[-\frac d{dt}+iT\sigma^{\mu\nu}(z^{\dag} F_{\mu\nu}z) e^{-\epsilon\frac d{dt}}+i\lambda e^{-\epsilon\frac d{dt}}\right] \nonumber\\ && \exp\left\{i\int^1_0\mathop{dt}\left[-\frac{\dot q^2}{4T}+iz^{\dag} \dot z+ z^{\dag} \dot q A(q) z- \lambda \right]\right\} \label{109} \end{eqnarray} But \begin{eqnarray} \lefteqn{\lim_{\epsilon \to +0} {\det}^{-1}\left[-\frac d{dt}+iT\sigma^{\mu\nu}(z^{\dag} F_{\mu\nu}z) e^{-\epsilon\frac d{dt}}+i\lambda e^{-\epsilon\frac d{dt}}\right]}\nonumber\\ &=&e^{-2i\int^1_0 \mathop{dt} \lambda (t)} {\det}^{-1}\left[-\frac d{dt}+iT\sigma^{\mu\nu}(z^{\dag} F_{\mu\nu}z) +i\lambda\right] \label{110} \end{eqnarray} \noindent by virtue of identity (\ref{66}). Further, \begin{eqnarray} \lefteqn{ {\det}^{-1}\left[-\frac d{dt}+iT\sigma^{\mu\nu}(z^{\dag} F_{\mu\nu}z) +i\lambda\right]} \nonumber\\ &=&{\det}^{-1}\left[-\gamma^0\frac d{dt}+\gamma^0iT\sigma^{\mu\nu}(z^{\dag} F_{\mu\nu}z) +i\gamma^0\lambda\right] \label{111} \end{eqnarray} \noindent The operator in R.H.S. of (\ref{111}) is {\it anti-Hermitean}. So we can write \begin{eqnarray} \lefteqn{ {\det}^{-1}\left[-\gamma^0\frac d{dt}+\gamma^0iT\sigma^{\mu\nu}(z^{\dag} F_{\mu\nu}z) +i\gamma^0\lambda\right]} \nonumber\\ &=&\int_{PBC}D^2\psi e^{i\int^1_0 \mathop{dt} (i\psi^{\dag} \gamma^0\dot{\psi}+T(\psi^{\dag}\gamma^0 \sigma{\mu\nu}\psi)(z^{\dag} F_{\mu\nu}z)+\lambda\psi^{\dag}\gamma^0\psi)} \label{112} \end{eqnarray} In the last formula $\psi^{\dag}$ and $\psi$ are already complex conjugate variables, measure $D^2\psi$ is defined by eq. (\ref{100.1}), and the "action" is real. So functional integral (\ref{112}) is well defined. Introducing standard notations $\bar{\psi}=\psi^{\dag}\gamma^0$ and substituting (\ref{110})-(\ref{112}) in (\ref{109}), one obtains, after integration with respect to $\lambda$, \begin{eqnarray} \lefteqn{ \mathop{{\rm tr}} e^{iT\wH}}\nonumber\\ &=&\dQM \nonumber\\ &&{\cal N}_T \exp\left\{i\int^1_0\mathop{dt}\left[-\frac{\dot q^2}{4T}+iz^{\dag} z +i\bar{\psi}\dot{\psi}+\right.\right.\nonumber\\ &&\left.\left.z^{\dag} \dot q A(q) z+ T(\bar{\psi}\sigma^{\mu\nu} \psi)(z^{\dag} F^{\mu\nu}(q)z)\right]\right\} \label{113} \end{eqnarray} The normalization constant ${\cal N}_T$ can be computed in the same way as its analog in eq. (\ref{99}): \begin{eqnarray} {\cal N}_T^{-1}&=&\left(- \frac{N}{2\pi^2T^2}\right)^{-1} \int\limits_{q(0)=q(1)=x} Dq\:\int_{PBC}D^2\psi D^2 z\delta\left( z^{\dag}z -1-\frac N 2 \right)\nonumber\\ &&\delta\left( \bar{\psi}\psi -3\right) \exp \left\{ i \int^1_0 \mathop{dt} \left[ -\frac{\dot q^2}{4T}+iz^{\dag} \dot z +i\psi^{\dag}\dot{\psi}\right]\right\} \label{114} \end{eqnarray} Substituting (\ref{113}) in (\ref{107}), we obtain desired representation for fer\-mi\-o\-nic determinant: {\samepage \begin{eqnarray} \lefteqn{ \ln{\det}(i\hat{\nabla}-m)}\nonumber\\ &=&\dTM\dQM\nonumber\\ &&{\cal N}_T \exp\left\{i \int^1_0\mathop{dt}\left[-\frac{\dot q^2}{4T}+iz^{\dag} z +i\bar{\psi}\dot{\psi}+\right.\right.\nonumber\\ &&\left.\left.z^{\dag} \dot q A(q) z+ T(\bar{\psi}\sigma^{\mu\nu} \psi)(z^{\dag} F^{\mu\nu}(q)z)\right]\right\} \label{115} \end{eqnarray} } The representation (\ref{115}) is manifestly gauge and Lorentz invariant and comprises only bosonic variables. The "action" in (\ref{115}) is real. So, as we will see soon, it is convenient for application of stationary phase method. Now let us derive bosonic path integral representation for generating functional $Z(j)$ of gauge invariant Green functions \begin{equation} G_{f_1,\ldots,f_n}(x_1,\ldots,x_n)= <T(\bar{\psi}_{f_1}(x_1)\psi_{f_1}(x_1)\ldots \bar{\psi}_{f_n}(x_n)\psi_{f_n}(x_n))> \label{116} \end{equation} The derivation is completely analogous to one given in the previous section for corresponding Euclidean correlators. For $Z(j)$ there exist standard path integral representation via anti-commuting variables: \begin{eqnarray} Z(j)&=&\int DA e^{iS_{YM}}\int D\bar{\Psi}D\Psi e^{ i\int dx\: \sum\limits_f \left[\bar{\Psi}_f(i\hat{\nabla}-m_f)\Psi_f+ j_f\bar{\Psi}_f\Psi_f\right]}\nonumber\\ &=& \int DA e^{iS_{YM}}\prod_f{\det}(i\hat{\nabla}-m_f+j_f) \label{117} \end{eqnarray} Repeating with minor changes the derivation of (\ref{106}), one gets {\samepage \begin{eqnarray} \lefteqn{ \ln{\det}(i\hat{\nabla}-m-j)}\nonumber\\ &=&\dTM\dQM\nonumber\\ &&{\cal N}_T \exp\left\{i\int^1_0\mathop{dt}\left[-\frac{\dot q^2}{4T}+iz^{\dag} z +i\bar{\psi}\dot{\psi}+ z^{\dag} \dot q A(q) z+ \right.\right.\nonumber\\ &&\left.\left.T(\bar{\psi}\sigma^{\mu\nu} \psi)(z^{\dag} F^{\mu\nu}(q)z)+ 2mTj(q)-iT\bar{\psi}\gamma^{\mu}\psi\partial_{\mu}j(q)- Tj^2(q) \right]\right\}\nonumber\\ &&\label{118} \end{eqnarray} 2} Substituting (\ref{118}) in (\ref{117}), we obtain worldline pat integral representation for $Z(j)$. Our next task is the investigation of quasiclassical approximation in QCD. To this end, we will formulate a scheme of evaluation of two-point function \begin{equation} G_{f_0}(x,y)=<T(\bar{\psi}_{f_0}(x)\psi_{f_0}(x)\bar{\psi}_{f_0}(y)\psi_{f_0}(y))> \label{119} \end{equation} This scheme can be easily generalized for evaluation of arbitrary Green functions. The equations for the stationary point will be interpreted as quasiclassical equations in QCD. They will be formulated in terms of particles that have spin and colour degrees of freedom and interecting with Yang-Mills field. First of all, we introduce more condenced notations: \begin{eqnarray} Q_f&\equiv& \{q_f,z^{\dag}_f,z_f,\bar{\psi}_f,\psi_f,T_f,m^2_f\} \label{120.0}\\ \int DQ_f(\cdots)&\equiv& -\frac 1 2 \int_0^{\infty} \frac{dT_f}{T_f} \: \int_{PBC} DqD^2\psi D^2 z\delta\left( z^{\dag}z -1-\frac N 2 \right)\nonumber\\ &&\delta\left( \bar{\psi}\psi -3\right) {\cal N}_T(\cdots) \label{122.1}\\ S[Q_f,A]&=& \int^1_0\mathop{dt}\left[-\frac{\dot q_f^2}{4T_f}+iz^{\dag}_f\dot z_f +i\bar{\psi_f}\dot{\psi_f}+ z^{\dag}_f \dot q A(q) z_f \right.\nonumber\\ &&\left.+ T_f(\bar{\psi_f}\sigma^{\mu\nu} \psi_f)(z^{\dag}_f F_{\mu\nu}(q)z_f)-T_fm^2_f\right] \label{120} \end{eqnarray} Further, using eq. (\ref{118}), one can get: \begin{equation} \frac{\delta}{\delta j(x)} \ln\det (i\hat{\nabla} -m_f+j) \raisebox{-10pt}{\rule{0.4pt}{20pt}${\scriptstyle j=0}$} =\int DQ_f e^{iS[Q_f,A]}R(x|Q_f) \label{121} \end{equation} {\samepage \begin{eqnarray} &&\frac{\delta^2}{\delta j(x)\delta j(y)} \ln\det (i\hat{\nabla} -m_f+j) \raisebox{-10pt}{\rule{0.4pt}{20pt}${\scriptstyle j=0}$} \nonumber\\ &=&\int DQ_f e^{iS[Q_f,A]}R(x|Q_f)R(y|Q_f)\nonumber\\ &+&i\delta(x-y)\int DQ_f e^{iS[Q_f,A]}T_f\int^1_0 dt_1dt_2\: \delta(q_{f}(t_1)-q_f(t_2)) \label{122} \end{eqnarray} } where \begin{equation} R(x|Q_f)= T\int^1_0 \mathop{dt}[2m_f-i\bar{\psi}_f(t)\gamma^{\mu}\psi_f(t)\partial_{\mu}] \delta(x-q_f(t)) \label{123} \end{equation} For any functional $W(j)$ \begin{equation} \frac{\delta^2 W(j)}{\delta j(x)\delta j(y)}= W(j)\left[ \frac{\delta^2 \ln W(j)}{\delta j(x)\delta j(y)}+ \frac{\delta \ln W(j)}{\delta j(x)} \frac{\delta \ln W(j)}{\delta j(y)} \right] \label{124} \end{equation} Using (\ref{117}), (\ref{121}), (\ref{122}) and applying (\ref{124}) for $W(j)=\det(i\hat{\nabla}-m_{f_0}+j)$, one obtains: \begin{equation} G_{f_0}(x-y)=G^{(1)}_{f_0}(x-y)+G^{(2)}_{f_0}(x-y) \label{125} \end{equation} \noindent where $G^{(1)}_{f_0}$ and $G^{(2)}_{f_0}$ correspond to the first and to the second terms in R.H.S. of (\ref{124}) respectively: \begin{eqnarray} G^{(1)}_{f_0}(x-y)&=& \int DADQ_{f_0}\: R(x,Q_{f_0})R(y,Q_{f_0}) \nonumber\\ &&\exp\left\{ iS_{YM}+iS[Q_{f_0},A] + \sum_f\int DQ_f \: e^{iS[Q_f,A]}\right\}\nonumber\\ &&+\delta(x-y)(\cdots) \label{126}\\ G^{(2)}_{f_0}(x-y)&=& \int DADQ_{f_0}D{Q'}_{f_0}\: [R(x,Q_{f_0})R(y,Q_{f_0})\nonumber\\ &&\exp\left\{\vphantom{ \sum_f\int DQ_f \: e^{iS[Q_f,A]}} iS_{YM}+iS[Q_{f_0},A]+iS[Q^{'}_{f_0},A]\right.\nonumber\\ &&\left.+ \sum_f\int DQ_f \: e^{iS[Q_f,A]}\right\} \label{127} \end{eqnarray} \noindent In (\ref{126}) $(\cdots)$ means the factor at $\delta(x-y)$ in R.H.S. of (\ref{122}). At first, we investigate the function $G^{(1)}_{f_0}$. The function $G_{f_0}(x-y)$, in itself, is defined up to counterterm $$const\, \delta(x-y)$$ \noindent by virtue of ultraviolet divergences. Then the last term in R.H.S. of (\ref{126}) only redefines this counterterm and so can be omitted. Expanding \begin{equation} \exp\left\{\sum_f\int DQ_f\: e^{iS[Q,A]}\right\} \label{128} \end{equation} \noindent in series, we can represent (\ref{126}) as \begin{eqnarray} \lefteqn{ G^{(1)}_{f_0}(x-y)}\nonumber\\ &=&\sum_{n_f}\frac 1{\prod_f n_f!} \int DADQ_{f_0}\: \left( \prod_f \prod_{j_f=1}^{n_f}DQ_{j_f}\right) R(x,Q_{f_0})R(y,Q_{f_0}) \nonumber\\ &&\exp\left\{iS_{YM}+iS[Q_{f_0},A]+ \sum_f\sum_{j_f=1}^{n_f} iS[Q_{j_f},A]\right\} \nonumber\\ &\equiv&\sum_{n_f}G^{(1)}_{f_0}(x-y;\{n_f\}) \label{129} \end{eqnarray} Obviously, each function $G^{(1)}_{f_0}(x-y;\{n_f\})$ correspond to contribution of all diagrams comprising $n_{f_1}$ quark loops of flavor $f_1$, $n_{f_2}$ quark loops of flavor $f_2$, etc. We will investigate each term in the series (\ref{129}) separately. At first we consider the term with $n_f=0$: \begin{equation} G^{(1)}_{f_0}(x-y;\{n_f=0\})= \int DADQ_{f_0}\: e^{iS_{YM}+iS[Q_{f_0},A]} R(x,Q_{f_0})R(y,Q_{f_0}) \label{130} \end{equation} The function $R(x|Q_f)$ can be represented as \begin{equation} R(x|Q_f)=\int d^4p\:\int^1_0\mathop{dt}\tilde R (p,t|Q_f)e^{ip(x-q_f(t))} \label{131} \end{equation} \noindent where \begin{equation} \tilde R (p,t|Q_f)=T(2m_f+p_{\mu}\bar{\psi}(t)\gamma^{\mu}\psi(t)) \label{132} \end{equation} \noindent Then we change variables: \begin{equation} q(t)\to q'(t)=q(t)-q_0, \ \ A(x)\to A(x-q_0) \label{132'} \end{equation} \noindent where the function $q'(t)$ obeys the boundary conditions \begin{equation} q'(0)=q'(1)=0 \label{132''} \end{equation} By virtue of translational invariance we get \begin{eqnarray} \lefteqn{ G^{(1)}_{f_0}(x-y;\{n_f=0\}) }\nonumber\\ &=&\int DA\int_{q(0)=q(1)=0}DQ_{f_0}\:\int dq_0\: e^{iS_{YM}+iS[Q_{f_0},A]}\nonumber\\ &&\int d^4p_1d^4p_2 \int^1_0 dt_1\:\int^1_0 dt_2\: \tilde R(p_1,t_1|q_{f_0}) R(p_2,t_2|q_{f_0})\nonumber\\ &&e^{ip_1(y-q_{f_0}(t_1))}e^{ip_2(y-q_{f_0}(t_2))} e^{-i(p_1+p_2)q_0}\nonumber\\ && \label{133} \end{eqnarray} Integration over $q_0$ gives $\delta(p_1+p_2)$, and we obtain the following representation for Fourier transformation of $G^{(1)}_{f_0}(x-y;\{n_f=0\})$: \begin{eqnarray} \lefteqn{ G^{(1)}_{f_0}(p;\{n_f=0\}) }\nonumber\\ &\equiv&\int dx\: e^{-ipx} G^{(1)}_{f_0}(x;\{n_f=0\}) \nonumber\\ &=&\int^1_0 {\mathop{dt}}_1 \int^1_0{\mathop{dt}}_2 \int DA\int_{q_{f_0}(0)= q_{f_0}(1)=0} DQ_{f_0}\: \tilde R (p,t_1|Q_{f_0})\tilde R(-p,t_2|Q_{f_0})\nonumber\\ &&\exp \left\{ iS_{YM}+iS[Q_{f_0},A]-ip(q_{f_0}(t_1)-q_{f_0}(t_2)) \right\} \label{134} \end{eqnarray} Now it is already easy to write equations for stationary point for the action in (\ref{134}). Omitting the index $f_0$, we get: \begin{eqnarray} \frac 1{g^2}\nabla_{\nu}F^{a\mu\nu}&=& \int^1_0 \mathop{dt} I^a(t)\dot q^{\mu}(t)\delta(x-q(t))\nonumber\\ &&+T\nabla_{\nu}\left[ \int^1_0 \mathop{dt} I^a(t) S^{\mu\nu} \delta(x-q(t))\right] \label{135} \end{eqnarray} \begin{equation} i\left( \frac d{dt} -i\dot q^{\mu}A_{\mu}(q) \right)z + TS^{\mu\nu}F_{\mu\nu}z=0 \label{136} \end{equation} \begin{equation} i\frac d{dt}\psi +TF^a_{\mu\nu}I^a\sigma^{\mu\nu}\psi=0 \label{137} \end{equation} \begin{equation} \frac 1 T \ddot q_{\mu}+\dot q^{\nu}F^a_{\mu\nu}I^a +\nabla_{\mu}F^a_{\nu\rho}(q)I^a S^{\nu\rho}=p_{\mu}(\delta(t_1)- \delta(t_2)) \label{138} \end{equation} \begin{equation} \frac 1{4T^2} \int^1_0 \mathop{dt} \dot q^2 +\frac 1 2 \int^1_0 \mathop{dt} S^{\mu\nu}F^a_{\mu\nu}I^a=m^2 \label{139} \end{equation} \noindent where \begin{equation} I^a=z^{\dag} \lambda^a z,\ \ S^{\mu\nu}=\bar{\psi}\sigma^{\mu\nu}\psi \label{140} \end{equation} Eqs. (\ref{135})-(\ref{138}) can be derived by variation of the action in (\ref{134}) with respect to $A^a_{\mu},\ z,\ \psi$, and $q^{\mu}$. In the derivation of (\ref{138}) we used (\ref{135})-(\ref{137}). The eq. (\ref{139}) is obtained by differentiation with respect to $T$. It easy to obtain closed system of equations in terms of $A^a_{\mu},\ z,\ I^a,$ and $S^{\mu\nu}$. Let \begin{equation} I=\lambda^aI^a \label{141} \end{equation} \noindent Then \begin{equation} iDI=TS^{\mu\nu}[F_{\mu\nu},I] \label{142} \end{equation} \begin{equation} \frac d{dt}S^{\mu\nu} =2TI^a F^{a[\mu}_{\ \ \ \ \rho}S^{\nu]\rho} \label{142'} \end{equation} \noindent where \begin{equation} DI\equiv \frac{dI}{dt}-i[\dot q^{\mu}A_{\mu}(q),I] \end{equation} Unknown functions in (\ref{135})-(\ref{138}) also obey boundary conditions \begin{equation} q(0)=q(1)=0,\ \ z(0)=z(1),\ \ \psi(0)=\psi(1) \label{144} \end{equation} \noindent They also satisfy the equations \begin{equation} z^{\dag} z=1+\frac N 2,\ \ \bar{\psi}\psi=3 \label{145} \end{equation} \noindent because of the presence of $\delta$-functions in the definition (\ref{122.1}) of the measure $DQ$. Apropos, we didn't introduce Lagrange multipliers to take into account these $\delta$-functions because the solutions of eqs. (\ref{136}), (\ref{137}) automatically satisfy the conditions \begin{equation} z^{\dag} z=const,\ \ \bar{\psi} \psi=const \label{146} \end{equation} Equations (\ref{135}), (\ref{138}), and (\ref{142}) are nothing but generalization of well-known Wong's equations \cite{24} that describe classical spinless particle interacting with Yang-Mills field. Indeed, if one omits the terms containing the tensor of spin $S^{\mu\nu}$ in eqs. (\ref{135}), (\ref{138}), and (\ref{142}) one gets just Wong's equations up to term \begin{equation} -p_{\mu}(\delta(t_1)-\delta(t_2)) \label{147} \end{equation} Eqs. (\ref{135})-(\ref{137}) admit simple interpretation. At "time" $t_1$ quark--anti-quark pair with momentum $p$ is created. Then quark and anti-quark move interacting with Yang-Mills field. The union of quark and anti-quark trajectories forms a closed loop passing the point $q=0$. (See (\ref{144})). At the "time" $t_2$ the quark--anti-quark pair annihilates. An analog of eqs. (\ref{135})-(\ref{139}) for terms with $n_f\ne 0$ in expansion (\ref{129}) can be derived in the similar way. Let $S^{\{n_f=0\}}$ be the action in the formula (\ref{134}) and $J^{a\mu}[Q]$ be the current in R.H.S. of eq. (\ref{135}). Then the analog of $S^{\{n_f=0\}}$ for the term with non zero numbers $\{n_f\}$ in (\ref{129}) is \begin{equation} S^{\{n_f\}}=S^{\{n_f=0\}}+\sum_f\sum_{j_f=1}^{n_f}S[Q_{j_f},A] \label{148} \end{equation} So instead of (\ref{135}) we have in this case equations \begin{equation} \frac 1{g^2}\nabla_{\nu}F^{a\mu\nu}=J^{a\mu}[Q_{f_0}]+ \sum_f\sum_{j_f=1}^{n_f}J^{a\mu}[Q_{j_f}] \label{150} \end{equation} Equations \begin{equation} \frac{\delta S^{\{n_f\}}}{\delta \psi_{j_f}}=0, \ \ \frac{\delta S^{\{n_f\}}}{\delta z_{j_f}}=0, \ \ \frac{\delta S^{\{n_f\}}}{\delta T_{f}}=0 \label{150'} \end{equation} \noindent coincide in form with (\ref{137}), (\ref{139}), and (\ref{140}), whereas the equation \begin{equation} \frac{\delta S^{\{n_f\}}}{\delta q_{j_f}}=0 \label{151} \end{equation} \noindent differs from (\ref{138}) by the term (\ref{147}). We see that quasiclassical configurations that give the main contribution in functional integrals (\ref{129}) are defined by equations of very similar structure. The same statement is valid for semiclassical configuration that give the main contribution in the function $G^{(2)}(x-y)$ as well as in any other gauge invariant Green functions. Therefore proposed scheme seems to be sufficiently general for application in QCD and other gauge theories. In this paper we restrict ourselves to formulation of general scheme of quasiclassical approximation in QCD leaving elaborating of details as well as applications for forthcoming papers. So at this point we stop our investigation of quasiclassical approximation in QCD. Brief discussion of possible applications will be given in the next section. \section{Conclusion} In the present paper we derived, first, new path integral representations for path ordered exponent (see eqs. (\ref{35})-(\ref{37.1}), (\ref{39}), (\ref{39.4}), (\ref{58})). We give two alternative, entirely independent derivation of these representations. So these results seems quite reliable. Then we applied these representations to derive new variant of non-Abelian Stokes theorem. Our result is represented by formula (\ref{75}). Then we transformed the latter to obtain another formulation of non-Abelian Stokes theorem that is similar to one proposed recently by Dyakonov and Petrov \cite{21}. As a result, we got corrected and generalized version of the theorem proved in \cite{21}. Dyakonov-Petrov version of non-Abelian Stokes theorem was already used in discussion of the role of monopoles in QCD \cite{21} and in attempts to derive string-like effective action in framework of QCD \cite{25}. So one can hope that our more general and simple version of this theorem will be also useful in discussion of various problems of QCD. Further, we derived pure bosonic worldline path integral representations for fermionic determinants as well as fermionic Green functions in Euclidean QCD. (See eqs. (\ref{99}), (\ref{105'}), and (\ref{106})). This representations comprise only integrations with respect to bosonic variables. On a finite lattice all integrals are quite simple and well convergent. (Remind, that domain of integration with respect to $z$ and $\psi$ in eqs. (\ref{99}), (\ref{105'}), and (\ref{106}) is bounded). So representation derived seem quite appropriate for lattice simulations. Our results for fermionic determinant and Green functions can be used also in another way. Namely, if one substitute instead of Wilson loop some phenomenological anzatz, one obtain formulation of the theory purely in terms of point particles. The most well-known example of such anzatz is Wilson area law: \begin{equation} <\mathop{\mbox{tr P}} \exp\left\{ i\oint \mathop{dx^{\mu}} A_{\mu}\right\}>= \exp\left\{ -KS_{min}+\left[{\mbox{perturbative}\atop \mbox{corrections}}\right]\right\} \label{152} \end{equation} This anzatz was applied, in particular, to derivation from QCD a quark--anti- quark potential used in potential models (see, for instance, recent papers \cite{26}, \cite{26.1} and references therein). Other anzatz for Wilson loops were proposed in framework of Dual QCD model \cite{27} and stochastic vacuum model \cite{28}. (See also \cite{26.1} for comparison of results obtained in framework of these models). Recently a sting-like expression in spirit of stochastic vacuum model was derived in papers \cite{25}, \cite{29}.Anzatz of another type, that gives an expression for Wilson loop in terms of trajectories of monopoles, was proposed in \cite{21}. All these results could be combined with ours to derive some effective action in terms of point particles that correctly describe colour and spin properties of quarks. Obviously, that such theory is much more simpler than initial quantum field theory and so thus approach of investigation seems to be rather perspective. Finally, in the present paper we also derived bosonic worldline path integral representation for fermionic determinants and Green functions in Minkowski space and started the investigation of the quasiclassical approximation in QCD. The key point in the formulation of quasiclassical approximation is quasiclassical QCD equations (\ref{135})-(\ref{139}) and (\ref{148})-(\ref{151}) which arise naturally when one applies the stationary point method to evaluation of functional integrals that defines Green functions in QCD. Quasiclassical equations derived appear to be nothing but a generalization of well-known Wong's equations. We formulated only those quasiclassical equations which arise in the problem of evaluation of the concrete Green function (\ref{119}). However, our method is quite general and one can easy to derive analogous equations in generic case. The next problem in investigation of quasiclassical approximation in QCD is the solution of quasiclassical equation of motion. Though these equations are very complicated, this problem doesn't seem hopeless, at least, in non-re\-la\-ti\-vis\-tic approximation. In electrodynamics in non-re\-la\-ti\-vis\-tic approximation one can neglect bremsstrahlung and retarded effects and, as a result, reformulate initial theory in terms of particles interacting by means of Coulomb forces. In the same way one may hope to derive potential of interaction of heavy quarks in QCD. Such approach is alternative to ones based on various anzatz for Wilson loops. Another interesting possibility to understand the quark confinement in framework of quasiclassical approach is connected with existence of classical solutions of Yang-Mills equations with singularity on the sphere. Such solution were discussed in the context of the problem of confinement recently in the papers \cite{29'} though the existence of such solutions was pointed out by several authors in 70's \cite{30}. Some solutions with singularity on the torus and cylinder was discussed in papers \cite{31}. If solutions with singularity on closed spacelike surface existed also for eqs. (\ref{135})-(\ref{139}) and their modifications, then quarks, moving inside such surface and interacting with Yang-Mills field, couldn't cross the surface. This would mean the confinement of quarks. So it is interesting to investigate singular solutions of equations (\ref{135})-(\ref{139}) and, if such solutions exist, to develop the corresponding quasiclassical perturbative theory. \section*{Appendix A} In this appendix we shall prove the validity of the identity (\ref{55}). \noindent We will prove (\ref{55}) by induction. For $N=2$ it is easy to prove (\ref{55}) by direct calculations because both sides of (\ref{55}) are simple rational functions of variables $\exp\{\frac i 2 \alpha_r\}$. Let us denote $$ x_r=e^{i\alpha_r} \eqno(A1) $$ \noindent Then L.H.S. of (\ref{55}) can be represented as $$ (-2i)^{N-1} \frac{ \prod\limits_{r=1}^N \sqrt x_r}{\prod\limits_{1\le p<q\le N}(x_p-x_q)} \sum_{j=1}^N(-1)^{j+1}x_j^N \prod_{{\scriptstyle {r,s\ne j\atop 1\le r<s\le N}}}(x_r-x_s) \eqno(A2) $$ \noindent whereas R.H.S. of (\ref{55}) as $$ (-2i)^{N-1}\left( \prod_{k=1}^N\sqrt x_k\right) \sum_{j=1}^N x_j \eqno(A3) $$ \noindent So eq. (\ref{55}) is equivalent to algebraic identity $$ \sum_{j=1}^N(-1)^{j+1}x_j^N \prod_{{\scriptstyle {r,s\ne j\atop 1\le r<s\le N}}}(x_r-x_s) = \left(\sum_{j=1}^N x_j\right) \prod\limits_{1\le r<s\le N}(x_r-x_s) \eqno(A4) $$ To prove (A4), we check, at first, that L.H.S. of (A4) is vanished if $x_i=x_j$. Obviously, it is sufficient to consider the case $x_1=x_2\equiv x$. If $x_1=x_2$ then only $j=1$ and $j=2$ terms survive in L.H.S. of (A4). The $j=1$ term is $$ x_1^N\prod\limits_{2\le r<s\le N}(x_r-x_s) = x^N\left(\prod_{p=3}^N(x-x_p) \right)\left(\prod\limits_{3\le p<q\le N}(x_p-x_q)\right) \eqno(A5) $$ \noindent whereas the $j=2$ term is equal to $$ -x_2^N \prod_{{\scriptstyle {r,s\ne 2\atop 1\le r<s\le N}}}(x_r-x_s) =- x^N\left(\prod_{p=3}^N(x-x_p) \right)\left(\prod\limits_{1\le p<q\le N}(x_p-x_q)\right) \eqno(A6) $$ \noindent So terms with $j=1$ and $j=2$ are cancelled. Thus L.H.S. of (A4) can be represented as $$ P(x_1,\ldots,x_N) \prod\limits_{1\le r<s\le N}(x_r-x_s) \eqno(A7) $$ \noindent We must prove that $$ P(x_1,\ldots,x_N) =\sum_{r=1}^N x_r \eqno(A8) $$ L.H.S. of eq. (A4) is polynomial of degree $N$ with respect to each variable $x_j$. Consequently, polynomial $P(x_1,\ldots,x_N)$ is linear in each $x_j$. So $$ P(x_1,\ldots,x_N) =a(x_2,\ldots,x_N)x_1+b(x_2,\ldots,x_n) \eqno(A9) $$ Comparing coefficient at $x_1^N$ in L.H.S. of (A4) and (A7), one find that $$a(x,\ldots,x_N)=1 \eqno(A10) $$ \noindent and so it remains to prove that $$b(x_2,\ldots,x_N)=\sum_{j=2}^N x_j \eqno(A11) $$ To prove (A11), let us compare the sums of $x_1$-independent terms in L.H.S. of (A4) and in (A7). They can be written as $$ \sum_{j=2}^N (-1)^{j+1} x_j^N \prod\limits_{{\scriptstyle {2\le r\le N \atop r\ne j}}} (-x_r) \prod\limits_{{\scriptstyle {2\le r<s\le N \atop r,s\ne j}}} (x_r-x_s) \eqno(A12)$$ \noindent and $$b(x_2,\ldots,x_N)\prod\limits_{2\le r\le N} (-x_r)\prod_{2\le r<s\le N}(x_r-x_s) \eqno(A13) $$ \noindent respectively. So (A11) is equivalent to $$ \sum_{j=2}^N(-1)^{j}x_j^{N-1} \prod_{{\scriptstyle {r,s\ne j\atop 2\le r<s\le N}}}(x_r-x_s) = \left(\sum_{j=2}^N x_j\right) \prod\limits_{2\le r<s\le N}(x_r-x_s) $$ But the latter equation is valid by virtue of induction assumption. Indeed, it can be obtained from (A4) by change $N\to N-1,\ ,x_1\to x_2, \ldots,x_{N-1}\to x_N$. This finishes the proof of identity (\ref{55}). \section*{Appendix B} In this appendix we prove the formula (\ref{66}). First, one notes that up to inessential factor $$\det\left( -\frac d{dt} + iBe^{-\epsilon\frac d{dt}}\right) =\det \oeo \eqno(B1) $$ \noindent So it is sufficient to prove that $$\lim_{\epsilon \to +0} \ln\det \oeo-\ln\det\owe= \frac i 2 \int^1_0\mathop{dt}\mathop{{\rm tr}} B \eqno(B2)$$ \noindent or, equivalently, $$ \lim_{\epsilon \to +0} \mathop{{\rm tr}}\ln\oeo-\mathop{{\rm tr}}\ln\owe=\frac i 2 \int^1_0\mathop{dt}\mathop{{\rm tr}} B \eqno(B3) $$ Let $\hat{\Pi}$ is orthogonal projector in $L^2(S^1)$ on any {\it finite} dimensional subspace. Then $$\lim_{\epsilon \to +0} \left[\mathop{{\rm tr}} \hat{\Pi} \ln\oeo -\mathop{{\rm tr}}\hat{\Pi}\ln\owe \right]=0 \eqno(B4)$$ \noindent In particular, let ${\cal H}_1$ be subspace in $L^2(S^1)$ that is orthogonal to one dimensional subspace spanned on the function $\phi_0(t)\equiv 1$. Then $$\lim_{\epsilon \to +0} \left[\mathop{{\rm tr}} \ln\oeo -\mathop{{\rm tr}}\ln\owe \right]$$ $$=\lim_{\epsilon \to +0} \left[{\mathop{{\rm tr}}}_{{\cal H}_1} \ln\oeo -{\mathop{{\rm tr}}}_{{\cal H}_1}\ln\owe \right] \eqno(B5)$$ In the space ${\cal H}_1$ the operator $$-\frac d{dt} e^{\epsilon\frac d{dt}} \eqno(B6)$$ \noindent is invertible. Let $G_{\epsilon}(t-t')$ is the kernel of $$\left(-\frac d{dt} e^{\epsilon\frac d{dt}}\right)^{-1} $$ The function $G_{\epsilon}(t-t')$ can be expresseed in terms of eigenfunctions $$\phi_n(t)=e^{2\pi int} ,\ \ n\ne 0 \eqno(B7)$$ \noindent of the operator (B6) and its eigenvalues $$\lambda_n=-2\pi in e^{2\pi in\epsilon},\ \ n\ne 0$$ Obviously, $$G_{\epsilon}(t-t')=\sum_{n\ne0}\frac{\phi_n(t)\bar{\phi_n}(t')} {\lambda_n} =-\frac 1{\pi}\sum_{n=1}^{\infty} \frac{\sin 2\pi n(t-t'-\epsilon)}{n} \eqno(B8)$$ Using the formula $$\sum_{n=1}^{\infty}\frac{\sin 2\pi n\epsilon}{\pi n}= \frac 1 2 -\epsilon \eqno(B9)$$ \noindent (that is valid for $0<\epsilon<1$), one finds that $$\lim_{\epsilon \to +0} G_{\epsilon}(0)=\frac 1 2 \eqno(B10)$$ \noindent whereas $$G_{0}(0)=0 \eqno(B10')$$ \noindent So $$G_{0}(0)\ne\lim_{\epsilon \to +0} G_{\epsilon}(0) \eqno(B11) $$ But for $-1<t<1,\ \ t\ne0$, $$\lim_{\epsilon \to +0} G_{\epsilon}(t)=G_{0}(t) \eqno(B11')$$ We will see soon that R.H.S. of (B2) is not vanished just by virtue of (B11). Further, up to inessential constant, $${\mathop{{\rm tr}}}_{{\cal H}_1}\ln\oeo={\mathop{{\rm tr}}}_{{\cal H}_1}\ln\left[ 1+i \left(-\frac d{dt} e^{\epsilon\frac d{dt}}\right)^{-1} B\right]$$ $$={\mathop{{\rm tr}}}_{{\cal H}_1}\left[ \left(-\frac d{dt} e^{\epsilon\frac d{dt}}\right)^{-1} B\right]+ \frac 1 2 {\mathop{{\rm tr}}}_{{\cal H}_1} \left(-\frac d{dt} e^{\epsilon\frac d{dt}}\right)^{-1} B \left(-\frac d{dt} e^{\epsilon\frac d{dt}}\right)^{-1} +\ldots \eqno(B12)$$ By virtue of equation $$\int^1_0 \mathop{dt} G_{\epsilon}(t-t')=0 \eqno(B13)$$ \noindent one can replace ${\mathop{{\rm tr}}}_{{\cal H}_1}$ by $\mathop{{\rm tr}}$ in all terms in the series (B12). So $${\mathop{{\rm tr}}}_{{\cal H}_1}\ln \oeo = i\int^1_0 \mathop{dt} G_{\epsilon}(0) \mathop{{\rm tr}} B(t)$$ $$+\frac 1 2 \int^1_0\mathop{dt}\int^1_0 dt'\: G_{\epsilon}(t-t') G_{\epsilon}(t'-t)\mathop{{\rm tr}} B(t)B(t')+\ldots \eqno(B14)$$ In the limit $\epsilon \to +0$ the first term in R.H.S. of (B14) is equal to $$\frac i 2 \int^1_0\mathop{dt} \mathop{{\rm tr}} B(t)$$ \noindent (see (B11)) whereas the sum of all others coincide with $${\mathop{{\rm tr}}}_{{\cal H}_1}\owe$$ \noindent by virtue of (B10'), (B11'). This proves the validity of the equation (B2). To check the formula (B2), let us consider the simplest case of one dimensional harmonic oscillator. Partion function $$Z(\beta)=\mathop{{\rm tr}}\exp\{-\beta a^{\dag} a\} \eqno(B15)$$ \noindent is known exactly: $$Z(\beta)=\sum_{n=0}^{\infty}e^{-\beta n}=\frac 1{1-e^{-\beta}} \eqno(B15)$$ On the other hand, $$Z(\beta)=\lim_{\epsilon \to +0}\int_{PBC}D\bar z Dz\: \exp\left\{\int^1_0\mathop{dt}(-\bar z\dot z -\beta \bar z e^{-\epsilon\frac d{dt}}z \right\}$$ $$={\det}^{-1}\left(-\frac d{dt}-\beta e^{-\epsilon\frac d{dt}} \right) \eqno(B16) $$ Putting in (B2) $B=i\beta$, one gets: $${\det}^{-1}\left(-\frac d{dt}-\beta e^{-\epsilon\frac d{dt}} \right) =e^{\frac{\beta}{2}} {\det}^{-1}\left(-\frac d{dt}-\beta \right) \eqno(B17)$$ The latter determinant we have already evaluated in the main text of the paper : $$ {\det}\left(-\frac d{dt}-\beta \right) =const \,\sinh \frac{\beta}2 \eqno(B18)$$ \noindent (see eq. (\ref{50}) for $N=1,\ \ \eta=\epsilon=0$). Comparing (B16)-(B19), we get $$Z(\beta)=const\,\frac{e^{\frac{\beta}2}}{\sinh\frac{\beta}2} \eqno(B19)$$ Evaluating the constant from the condition $$Z(\infty)=1$$ \noindent we reproduce the true answer (B15). We would like to stress that if we had put $\epsilon=0$ in (B16) before evaluating of the functional integral the result would have been wrong.
proofpile-arXiv_065-431
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} \setcounter{footnote}{0} A still challenging question in strong interaction physics is the derivation of the low-energy properties of the spectrum from ``QCD first principle", due to our limited present skill with non-perturbative physics. At very low energy, where the ordinary perturbation theory cannot be applied, Chiral Perturbation Theory~\cite{GaLeut} and Extended Nambu--Jona-Lasinio models~\cite{NJL,ENJL}~\footnote{For a recent complete review see \cite{Miransky}.} give a consistent framework in terms of a set of parameters that have to be fixed from the data; yet the bridge between those effective parameters and the basic QCD degrees of freedom remains largely unsolved. Although lattice QCD simulations recently made definite progress~\cite{lattcsb} in that direction, the consistent treatment of dynamical unquenched quarks and the chiral symmetry remains a serious problem. \\ In this paper, we investigate a new, semi-analytical method, to explore {\it how far} the basic QCD Lagrangian can provide, in a self-consistent way, non-zero dynamical quark masses, quark condensates, and pion decay constant, in the limit of vanishing Lagrangian (current) quark masses. Such a qualitative picture of chiral symmetry breakdown (CSB) can be made more quantitative by applying a new ``variational mass" approach, recently developed within the framework of the anharmonic oscillator \cite{bgn}, and in the Gross--Neveu (GN) model \cite{gn1,gn2}. The starting point is very similar to the ideas developed a long time ago and implemented in various different forms in refs.\cite{pms}, \cite{delta}. There, it was advocated that the convergence of conventional perturbation theory may be improved by a variational procedure in which the separation of the action into ``free" and ``interaction" parts is made to depend on some set of auxiliary parameters. The results obtained by expanding to finite order in this redefined perturbation series are optimal in regions of the space of auxiliary parameters where they are least sensitive to these parameters. Recently there appeared strong evidence that this optimized perturbation theory may indeed lead to a rigorously convergent series of approximations even in strong coupling cases \cite{JONE}. \\ An essential novelty, however, in \cite{bgn}--\cite{gn2} and the present paper, is that our construction combines in a specific manner the renormalization group (RG) invariance with the properties of an analytically continued, arbitrary mass parameter $m$. This, at least in a certain approximation to be motivated, allows us to reach {\it infinite} order of the variational-perturbative expansion, and therefore presumably optimal, provided it converges. Our main results are a set of non-perturbative ansatzs for the relevant CSB quantities, as functions of the variational mass $m$, which can be studied for extrema and optimized. Quite essentially, our construction also provides a simple and consistent treatment of the renormalization, reconciling the variational approach with the inherent infinities of quantum field theory and the RG properties. Before proceeding, let us note that there exists a quite radically different attitude towards CSB in QCD, advocating that the responsible mechanism is most probably the non-perturbative effects due to the {\em instanton} vacuum~\cite{Callan}, or even more directly related to confinement~\cite{Cornwall}. However, even if the instanton picture of CSB is on general grounds well motivated, and many fruitful ideas have been developed in that context\footnote{See e.g ref.~\cite{Shuryak} for a review and original references.}, as far as we are aware there is at present no sufficiently rigorous or compelling evidence for it, derived from ``first principle". In any event, it is certainly of interest to investigate quantitatively the ``non-instantonic" contribution to CSB, and we hope that our method is a more consistent step in that direction. \section{Dynamical quark masses} In what follows we only consider the $SU(n_f)_L \times SU(n_f)_R$ part of the chiral symmetry, realized by the QCD Lagrangian in the absence of quark mass terms, and for $n_f =2$ or $n_f = 3$ as physically relevant applications. Following the treatment of the anharmonic oscillator~\cite{bgn} and its generalization to the GN model~\cite{gn1,gn2}, let us consider the following modification of the usual QCD Lagrangian, \begin{equation} L_{QCD} \to L^{free}_{QCD}(g_0 =0, m_0=0) -m_0 \sum^{n_f}_{i=1} \overline{q}_i q_i + L^{int}_{QCD}(g^2_0 \to x g^2_0) +x \; m_0 \sum^{n_f}_{i=1} \overline{q}_i q_i\;, \label{xdef} \end{equation} where $L^{int}_{QCD}$ designates the ordinary QCD interaction terms, and $x$ is a convenient ``interpolating" expansion parameter. This formally is equivalent to substituting everywhere in the bare Lagrangian, \begin{equation} m_0 \to m_0\; (1-x); ~~~~g^2_0 \to g^2_0\; x, \label{substitution} \end{equation} and therefore in any perturbative (bare) quantity as well, calculated in terms of $m_0$ and $g^2_0$. Since the original massless QCD Lagrangian is recovered for $x \to 1$, $m_0$ is to be considered as an {\it arbitrary } mass parameter after substitution (\ref{substitution}). One expects to optimize physical quantities with respect to $m_0$ at different, possibly arbitrary orders of the expansion parameter $x$, eventually approaching a stable limit, i.e {\it flattest} with respect to $m_0$, at sufficiently high order in $x$. \\ However, before accessing any physical quantity of interest for such an optimization, the theory should be renormalized, and there is an unavoidable mismatch between the expansion in $x$, as introduced above, and the ordinary perturbative expansion as dictated by the mass and coupling counterterms. Moreover, it is easy to see that at any finite order in the $x$ expansion, one always recovers a trivial result in the limit $ m \to 0$ (equivalently $x\to 1$), which is the limit in which to identify non-zero order parameters of CSB. These problems can be circumvented by advocating a specific ansatz, which resums the (reorganized) perturbation series in $x$ and is such that the limit $x \to 1$ no longer gives a trivial zero mass gap. As was shown in detail in ref.~\cite{gn2}, the ansatz for the dynamical mass is most easily derived by following the steps\footnote{See also ref.~\cite{jlk} for a detailed derivation in the QCD context.}:\\ {\it i}) Consider first the general solution for the running mass, given as \begin{equation} m(\mu^{'}) = m(\mu )\;\; {\exp\left\{ -\int^{g(\mu^{'})}_{g(\mu )} dg {\gamma_m(g) \over {\beta(g)}} \right\} } \label{runmass} \end{equation} in terms of the effective coupling $g(\mu)$, whose RG evolution is given as $\mu dg(\mu)/d\mu \equiv \beta(g)$, and $\gamma_m(g) \equiv -(\mu/m)d(m(\mu))/d\mu$. Solving (\ref{runmass}) imposing the ``fixed point" boundary condition: \begin{equation} M \equiv m(M), \label{RGBC} \end{equation} at two-loop RG order we obtain, after some algebra (we use the normalization $\beta(g) = -b_0 g^3 -b_1 g^5 -\cdots$, $\gamma_m(g) = \gamma_0 g^2 +\gamma_1 g^4 +\cdots$): \begin{equation} M_2 = \bar m \;\; \displaystyle{f^{-\frac{ \gamma_0}{2b_0}}\;\; \Bigl[\frac{ 1 +\frac{b_1}{b_0} \bar g^2 f^{-1}}{ 1+\frac{b_1}{b_0}\bar g^2} \Bigr]^{ -\frac{\gamma_1}{ 2 b_1} +\frac{\gamma_0}{2 b_0} } }\;, \label{MRG2} \end{equation} where $\bar m \equiv m(\bar\mu)$, $\bar g \equiv g(\bar\mu)$ ($\bar \mu \equiv \mu \sqrt{4 \pi} e^{-\gamma_E/2}$), and $f \equiv \bar g^2/g^2(M_2)$ satisfies \begin{equation} f = \displaystyle{ 1 +2b_0 \bar g^2 \ln \frac{M_2}{\bar \mu } +\frac{b_1}{b_0} \bar g^2 \ln \Bigl[\frac{ 1 +\frac{b_1}{b_0} \bar g^2 f^{-1}}{ 1 +\frac{b_1}{b_0} \bar g^2 }\;f\;\Bigr] }\; ; \label{f2def} \end{equation} (note in (\ref{MRG2}) and (\ref{f2def}) the recursivity in both $f$ and $M_2$). The necessary non-logarithmic perturbative corrections to those pure RG results are then consistently included as \begin{equation} M^P_2 \equiv M_2 \;\Bigl(1 +{2\over 3}\gamma_0 {\bar g^2\over f} +{K \over{(4 \pi^2)^2}}{\bar g^4\over f^2}+{\cal O}(g^6)\;\Bigr)\;, \label{Mpole} \end{equation} where the complicated two-loop coefficient $K$ was calculated exactly in ref.~\cite{Gray}. Equation.~(\ref{Mpole}) defines the (infrared-convergent, gauge-invariant) pole mass~\cite{Tarrach} $M^P_2$, in terms of the $\overline{MS}$ mass at two-loop order, and can be shown~\cite{jlk} to resum the leading (LL) {\it and} next-to-leading logarithmic (NLL) dependence in $\bar m$ to all orders. \\ {\it ii}) Perform in expressions (\ref{MRG2}), (\ref{f2def}), (\ref{Mpole}) the substitution $ \bar m \to \bar m v $, and integrate the resulting expression, denoted by $M^P_2(v)$, according to \begin{equation} \frac{1}{2i\pi} \;\oint \frac{dv}{v}\;e^v M^P_2(v)\;, \label{contgen} \end{equation} where the contour is around the negative real $v$ axis. \\ In \cite{gn2} it was shown that the previous steps correspond (up to a specific renormalization scheme (RS) change, allowed on general grounds from RG properties) to a resummation of the $x$ series as generated from the substitution (\ref{substitution})\footnote{$v$ is related to the original expansion parameter $x$ as $x = 1-v/q$, $q$ being the order of the expansion.}. Moreover this is in fact the only way of rendering compatible the above $x$ expansion and the ordinary perturbative one, thus obtaining finite results. Actually the resummation coincides with the exact result in the large-$N$ limit of the GN model. Now, since the summation can be formally extended to arbitrary RG orders~\cite{gn2}, including consistently as many arbitrary perturbative correction terms as known in a given theory, in the QCD case we make the assumption that it gives an adequate ``trial ansatz", to be subsequently optimized in a way to be specified next. After appropriate rescaling of the basic parameters, $\bar g$ and $\bar m$, by introducing the RG-invariant basic scale $\Lambda_{\overline{MS}}$~ \cite{Lambda} (at two-loop order), and the convenient scale-invariant dimensionless ``mass" parameter \begin{equation} m''\equiv \displaystyle{(\frac{\bar m}{ \Lambda_{\overline{MS}}}) \; 2^{C}\;[2b_0 \bar g^2]^{-\gamma_0/(2b_0)} \;\left[1+\frac{b_1}{b_0}\bar g^2\right]^{ \gamma_0/(2 b_0)-\gamma_1/(2 b_1)}} \; , \label{msec2def} \end{equation} we end up with the following dynamical mass ansatz: \begin{equation} { M^P_2 (m^{''})\over \Lambda_{\overline{MS}}} = {2^{-C} \over{2 i \pi}} \oint dy {e^{\;y/m^{''}} \over{F^A(y) [C + F(y)]^B}} {\left(1 +{{\cal M}_{1}\over{F(y)}} +{{\cal M}_{2}\over{F^2(y)}} \right)}, \label{contour7} \end{equation} where $y \equiv m'' v$, and $F$ is defined as \begin{equation} F(y) \equiv \ln [y] -A \; \ln [F(y)] -(B-C)\; \ln [C +F(y)], \label{Fdef} \end{equation} with $A =\gamma_1/(2 b_1)$, $B =\gamma_0/(2 b_0)-\gamma_1/(2 b_1)$, $C = b_1/(2b^2_0)$, in terms of the RG coefficients \cite{betagamma}. Finally the perturbative corrections in (\ref{contour7}) are simply given as ${\cal M}_{1} =(2/3)(\gamma_0/2b_0)$ and ${\cal M}_{2} = K/(2b_0)^2$. \\ Observe in fact that, were we in a simplified QCD world, where there would be {\em no} non-logarithmic perturbative contributions (i.e. such that ${\cal M}_{1} = {\cal M}_{2} = \cdots = 0$ in (\ref{contour7})), the latter ansatz would then resums exactly the $x$ variational expansion. In that case, (\ref{contour7}) would have a very simple behaviour near the origin $m'' \to 0$. Indeed, it is easy to see that (\ref{Fdef}) admits an expansion $ F(y) \simeq C^{(B-C)/A}\;y^{1/A}$ for $y \to 0$, which immediately implies that (\ref{contour7}) would give a simple pole at $y \to 0$, with a residue giving $M_2 = (2C)^{-C}\;\Lambda_{\overline{MS}} $. Moreover one can always choose an appropriate renormalization scheme in which $ b_2$ and $\gamma_2$ are set to zero, as well as all higher order coefficients, so that there are no other corrections to the simple above relation. Now, in the realistic world, ${\cal M}_1$, ${\cal M}_2$, etc can presumably not be neglected. We can nevertheless expand (\ref{contour7}) near $m'' \to 0$ for any known non-zero ${\cal M}_{i}$, using \begin{equation} \label{hankel} \frac{1}{2i \pi} \oint dy e^{y/m^{''}} y^\alpha = \frac{(m^{''})^{1+\alpha}}{\Gamma[-\alpha]}\; , \end{equation} and the resulting Laurent expansion in $(m'')^{1/A}$ may be analysed for extrema and optimized at different, in principle arbitrary $(m'')^{1/A}$ orders. An important point, however, is that the perturbative corrections do depend on the RS choice, as is well known. Since the pure RG behaviour in (\ref{contour7}) already gives the order of magnitude, $M \simeq {\rm const} \times \Lambda_{\overline{MS}}$, we can hope that a perturbative but optimized treatment of the remaining corrections is justifed. In other words we shall perform an ``optimized perturbation" with respect to $m''$ around the non-trivial fixed point of the RG solution. \\ To take into account this RS freedom, we first introduce in (\ref{contour7}) an arbitrary scale parameter $a$, from $\bar \mu \to a\; \bar \mu$. Accordingly the perturbative coefficients ${\cal M}_{i}$ in (\ref{contour7}) take a logarithmic dependence in $a$, simply fixed order by order from (\ref{MRG2})--(\ref{Mpole}) and the requirement that (\ref{contour7}) differs from the original $\overline{MS}$ expression only by higher order terms. The $a$-dependence will eventually exhibit a non-trivial extrema structure and we shall also optimize the result with respect to $a$\footnote{This procedure indeed gave very good results~\cite{gn2} in the GN model, where in particular for low values of $N$ the optimal values found, $a_{opt}$, are quite different from 1.}. Actually there are other possible changes of renormalization prescriptions affecting expression (\ref{contour7}) in addition to the $a$ dependence, which may be taken into account as well. More precisely, the second coefficient of $\gamma_m(g)$, $\gamma_1$, do depend on the RS choice, even in MS schemes~\cite{Collins}. As it turns out, this additional RS freedom is very welcome in our case: in fact, the previous picture is invalidated, due to the occurence of extra branch cuts in the $y$ plane at $Re[y_{cut}] > 0$, as given by the zeros of $dy/dF$ from (\ref{Fdef}) (in addition to the original cut on the negative real $y$ axis). This prevents using the expansion near the origin, eq.~(\ref{hankel}), since it would lead to ambiguities of ${\cal O}(\exp(Re[y]/m''))$ for $m'' \to 0$\footnote{ The origin of those singularities is rather similar to the ambiguities related to renormalons~\cite{renormalons}. An essential difference, however, is that the present singularities occur in the analytic continuation of a mass parameter rather than a coupling constant, and it is possible to move those singularities away by an appropriate RS change, as we discuss next. See ref.~\cite{jlk} for an extended discussion.}. The specific contour around the negative real axis was suggested by the known properties of the large $N$ limit of the GN model, and it is not surprising if the analytic structure is more complicated in QCD. However, the nice point is that the actual position of those cuts do depend on the RS, via $A(\gamma_1)$ in (\ref{Fdef}). Defining $ \gamma^{'}_1 \equiv \gamma_1 +\Delta \gamma_1$, we can choose $Re[y_{cut}] \simeq 0$ for $\Delta\gamma_1 \simeq$ 0.00437 (0.00267) for $n_f =$ 2 ($n_f =$ 3), respectively. We therefore consider~\cite{jlk} the general RS change \begin{equation} m' = \bar m\;(1+B_1 \bar g^2 +B_2 \bar g^4)\;;\;\;\;g^{'2} = \bar g^2\; (1 +A_1 \bar g^2 +A_2 \bar g^4)\; \label{RSchange} \end{equation} (implying $\Delta\gamma_1 = 2b_0 B_1 -\gamma_0 A_1$), and optimize with respect to this new arbitrariness\footnote{ We also impose a further RS choice, $ b^{'}_2 = 0$, $\gamma^{'}_2 = 0$, which fixes $A_2$, $B_2$ in (\ref{RSchange}) and guarantees that our two-loop convention for $\Lambda_{\overline{MS}}$ remains unaffected. Note, however, that (\ref{RSchange}) implies $\Lambda_{\overline{MS}} \to \Lambda_{\overline{MS}} \; \exp\{\frac{A_1}{2b_0}\}\equiv \Lambda'$. In what follows we express the results in terms of the original $\Lambda_{\overline{MS}}$.}. However one soon realizes that our extension of the ``principle of minimal sensitivity" (PMS)~\cite{pms} defines a rather complicated optimization problem. Fortunately, we can study this problem within some approximations, which we believe are legitimate. Since the ansatz (\ref{contour7}) (with the above RS change understood, to make it consistent) would indeed be optimal with respect to $m^{''}$ for {\em vanishing} perturbative non-logarithmic corrections, ${\cal M}_{i} =0$, we shall assume that the expansion for small $m''$ is as close as possible to an optimum, and define the $m^{''} \to 0$ limit by some relatively crude but standard approximation, avoiding numerical optimization with respect to $m^{''}$. The approximation we are looking for is not unique: given (\ref{contour7}), one could construct different approximations leading to a finite limit for $m'' \to 0$~\cite{gn2}. Here we shall only demonstrate the feasibility of our program in the simplest possible realization. In fact, since we shall anyhow optimize with respect to the RS dependence we assume that it largely takes into account this non-uniqueness due to higher order uncertainties. Pad\'e approximants are known to greatly improve perturbative results~\cite{pade} and often have the effect of smoothing the RS dependence. We thus take a simple Pad\'e approximant which by construction restitutes a simple pole for $F \to 0$ (i.e $m'' \to 0$) in (\ref{contour7}), and gives \begin{equation} {M^{Pad\acute{e}}(a,\Delta\gamma_1,B_1,m''\to 0) = \Lambda_{\overline{MS}}\; (2C)^{-C} \;a\; \exp\{\frac{A_1}{2b_0}\}\;\left[ 1 -{{\cal M}^2_{1}(a, \Delta\gamma_1, B_1)\over{{\cal M}_{2}(a, \Delta\gamma_1, B_1)}}\right] } \label{Mpade} \end{equation} We have performed a rather systematic study of the possible extrema of (\ref{Mpade}) for arbitrary $a$, $B_1$ (with $\Delta\gamma_1$ fixed so that the extra cuts start at $ Re[y] \simeq 0 $). We obtain the flattest such extrema for $a \simeq 2.1$, $B_1 \simeq 0.1$, which leads to the result \begin{equation} M^{Pad\acute{e}}_{opt}(m''\to 0) \simeq 2.97\;\Lambda_{\overline{MS}}(2)\; \label{Mnum} \end{equation} for $n_f=2$. Similarly, we obtain $M^{Pad\acute{e}}_{opt}(m''\to 0) \simeq 2.85 \Lambda_{\overline{MS}}(3)$ for $n_f=3$. Note that these values of the dynamical quark masses, if they are to be consistent with the expected range~\cite{Miransky} of $M_{dyn}\simeq$ 300-400 GeV, call for relatively low $\Lambda_{\overline{MS}} \simeq $ 100-150 GeV, which is indeed supported by our results in the next section. \section{Composite operators and $F_\pi$} We shall now generalize the ansatz (\ref{contour7}) for the pion decay constant $F_\pi$. The main idea is to do perturbation theory around the same RG evolution solution with the non-trivial fixed point, as specified by the function $F$ in (\ref{Fdef}), with perturbative correction terms obviously specific to $F_\pi$. A definition of $F_\pi$ suiting all our purposes is~\cite{GaLeut,derafael} \begin{equation} i\;\int d^4q e^{iq.x} \langle 0 \vert T\;A^i_\mu(x) A^k_\nu(0) \vert 0 \rangle = \delta^{ik} g_{\mu \nu} F^2_\pi +{\cal O}(p_\mu p_\nu) \label{Fpidef} \end{equation} where the axial vector current $A^i_\mu \equiv (\bar q \gamma_\mu \gamma_5\lambda^i q)/2$ (the $\lambda^i$'s are Gell-Mann $SU(3)$ matrices or Pauli matrices for $n_f =3$, $n_f=2$, respectively). Note that according to (\ref{Fpidef}), $F_\pi$ is to be considered as an order parameter of CSB~\cite{Stern}. \\ The perturbative expansion of (\ref{Fpidef}) for $m \neq 0$ is available to the three-loop order, as it can be easily extracted from the very similar contributions to the electroweak $\rho$-parameter, calculated at two loops in \cite{abdel} and three loops in \cite{Avdeev}. The appropriate generalization of (\ref{contour7}) for $F_\pi$ now takes the form \begin{eqnarray} & \displaystyle{{F^2_\pi \over{\Lambda_{\overline{MS}}^2}} = (2b_0)\; {2^{-2 C} a^2\over{2 i \pi}} \oint {dy\over y}\; y^2 {e^{y/m^{''}}} \frac{1}{F^{\;2 A-1} [C + F]^{\;2 B}} } \; \times \nonumber \\ & \displaystyle{ {\delta_{\pi } \left(1 +{\alpha_{\pi}(a)\over{F}}+{\beta_{\pi}(a) \over{F^2}} \;+\cdots \right)} } \label{Fpiansatz} \end{eqnarray} in terms of $F(y)$ defined by eq.~(\ref{Fdef}) and where $\delta_\pi$, $\alpha_\pi(1)$ and $\beta_\pi(1)$, whose complicated expressions will be given elsewhere~\cite{jlk}, are fixed by matching the perturbative $\overline{MS}$ expansion in a way to be specified next. Formula (\ref{Fpiansatz}) necessitates some comments: apart from the obvious changes in the powers of $F$, $y$, etc, dictated by dimensional analysis, note that the perturbative expansion of the (composite operator) $\langle A_\mu A_\nu \rangle$ in (\ref{Fpidef}) starts at one-loop, but zero $g^2$ order. This leads to the extra $2b_0 F$ factor in (\ref{Fpiansatz}), corresponding to an expansion starting at ${\cal O}(1/g^2)$\footnote{ The ${\cal O}(1/g^2)$ first-order term cancels anyhow after the necessary subtraction discussed below.}. Another difference is that the perturbative expansion of (\ref{Fpidef}) is ambiguous due to remaining divergences after mass and coupling constant renormalization. Accordingly it necessitates additional subtractions which, within our framework, are nothing but the usual renormalization procedure for a composite operator, which is (perturbatively) well-defined~\cite{Collins}. The only consequence is that, after a consistent treatment of the subtracted terms (i.e respecting RG invariance), the unambiguous determination of the $1/F^n$ perturbative terms in (\ref{Fpiansatz}) necessitates the knowledge of the $(n+1)$ order of the ordinary perturbative expansion. The nice thing, however, is that the subtracted terms only affect the values of $\alpha_\pi$ and $\beta_\pi$, but not the {\em form} of the ansatz (\ref{Fpiansatz}), as soon as the order of the variational-perturbative expansion is larger than 1~\cite{gn2}. The consistency of our formalism is checked by noting that the re-expansion of (\ref{Fpiansatz}) do reproduce correctly the LL and NLL dependence in $\bar m$ of the perturbative expansion of the composite operator to all orders. The analyticity range with respect to $\Delta\gamma_1$, discussed in section 2, remains valid for (\ref{Fpiansatz}) as well, since the branch cuts are determined by the very same relation (\ref{Fdef}). We can thus proceed to a numerical optimization with respect to the RS dependence, along the same line as the mass case in section 2. Using an appropriate Pad\'e approximant form to define the $F \to 0$ ($m'' \to 0$) limit, we obtain the optimal values as \begin{equation} F^{Pad\acute{e}}_{\pi ,opt}(m'' \to 0) \simeq 0.55\;\Lambda_{\overline{MS}}(2)\;\;\;(0.59\;\Lambda_{\overline{MS}}(3)\;)\;, \end{equation} for $n_f =$ 2 (3). With $F_\pi \simeq 92$ MeV, this gives $\Lambda_{\overline{MS}} \simeq $ 157 (168) MeV, for $n_f =$ 3 (2). \section{$\langle \bar q q \rangle$ ansatz} As is well known~\cite{Collins,Miransky}, $\langle \bar q q \rangle$ is not RG-invariant, while $m \langle \bar q q \rangle$ is; this is thus the relevant quantity to consider for applying our RG-invariant construction. A straightforward generalization of the derivation in section 3 leads to the ansatz \begin{eqnarray} {\bar m \langle \bar q q\rangle \over{\Lambda_{\overline{MS}}^4}} =(2b_0) {2^{-4 C} a^4\over{2 i \pi}} \oint {dy\over y} {e^{y/m^{''}} y^4 \over{(F)^{4 A-1} [C + F]^{4 B }}} { \delta_{\langle \bar q q\rangle} \left(1 +{\alpha_{\langle \bar q q\rangle}(a)\over{F(y)}} \;\right)} \label{qqansatz} \end{eqnarray} where again the coefficients $\delta_{\langle\bar q q\rangle}$ and $\alpha_{\langle\bar q q\rangle}(1)$ are obtained from matching the ordinary perturbative expansion after a subtraction, and will be given explicitely elsewhere~\cite{jlk}. The perturbative expansion, known up to two-loop order \cite{Spiridonov,jlk} implies that one only knows unambiguously the first order correction ${\cal O}(1/F)$ in (\ref{qqansatz}), as previously discussed. Apart from that, (\ref{qqansatz}) has all the expected properties (RG invariance, resumming LL and NLL dependence etc), but a clear inconvenience is that $\langle\bar q q\rangle$ cannot be directly accessed, being screened by tiny explicit symmetry breaking effects due to $m \neq 0$. This is of course a well-known problem, not specific to our construction. However, it is not clear how to consistently include explicit symmetry breaking effects within our framework. As amply discussed, in (\ref{qqansatz}) $m^{''}$ is an arbitrary parameter, destined to reach its chiral limit $m^{''} \to 0$. Accordingly, $\bar m \to 0$ for $m'' \to 0$, so that one presumably expects only to recover a trivial result, $\bar m \langle\bar q q\rangle \to 0$ for $m'' \to 0$. This is actually the case: although (\ref{qqansatz}) potentially gives a non-trivial result in the chiral limit, namely the simple pole residue ($\equiv 2b_0(2C)^{-C} \;\delta_{\langle \bar q q\rangle} \;\alpha_{\langle \bar q q\rangle}(a)$, upon neglecting unknown higher-order purely perturbative corrections), when we require extrema of this expression with respect to RS changes, using for the $m'' \to 0$ limit a Pad\'e approximant similar to the one for $F_\pi$, we do {\em not} find non-zero extrema. Such a result is not conclusive regarding the actual value of $\langle\bar q q \rangle(\bar\mu)$, but it may be considered a consistency cross-check of our formalism. \\ On the other hand, we should mention that our basic expression (\ref{qqansatz}) {\em does} possess non-trivial extrema for some $m''_{opt} \neq 0$. These we however refrain from interpreating in a more quantitative way since, within our framework, we cannot give to $m''\times \Lambda_{\overline{MS}}$ the meaning of a true, explicit quark mass (whose input we in principle need in order to extract a $\langle \bar q q\rangle $ value from (\ref{qqansatz})). At least, it strongly indicates that it should be possible to extract $\langle \bar q q\rangle$ in the chiral limit, by introducing in a consistent way a small explicit symmetry-breaking mass, $-m_{0,exp} \bar q_i q_i$, to the basic Lagrangian (\ref{xdef}). \section{Summary} In this paper we have shown that the variational expansion in arbitrary $m''$, as developed in the context of the GN model~\cite{gn2}, can be formally extended to the QCD case, apart from the complication due to the presence of extra singularities, which can be however removed by appropriate RS change. As a result we obtain in the chiral limit non-trivial relationships between $\Lambda_{\overline{MS}}$ and the dynamical masses and order parameters, $F_\pi$, $\bar m \langle\bar q q\rangle$. The resulting expressions in a generalized RS have been numerically optimized, using a well-motivated Pad\'e approximant form, due to the complexity of the full optimization problem. The optimal values obtained for $M_q$ and $F_\pi$ are quite encouraging, while for $\langle\bar q q\rangle$ they are quantitatively not conclusive, due to the inherent screening by an explicit mass term of this quantity, in the limit $m \to 0$. A possible extension to include consistently explicit breaking mass terms in our formalism is explored in ref.~\cite{jlk}. \vskip .5 cm {\large \bf Acknowledgements} \\ We are grateful to Eduardo de Rafael for valuable remarks and discussions. J.-L. K. also thanks Georges Grunberg, Heinrich Leutwyler, Jan Stern and Christof Wetterich for useful discussions. (C.A) is grateful to the theory group of Imperial College for their hospitality.\\
proofpile-arXiv_065-432
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{#1}\setcounter{equation}{0}} \newcommand\Sect[2] {\sect{#1}\label{s.#2} \ifnum0=1 \query{s.#2} \fi} \def\semitimes {\begin{picture}(9.4,8)\put(1.57,0.32){\line(0,1) {4.4}}\put(0,0){$\times$} \end{picture} } \def\mbox{$^{\scriptscriptstyle(h)}\!S$} {\mbox{$^{\scriptscriptstyle(h)}\!S$}} \def\mbox{$^{\scriptscriptstyle(hh')}\!S$} {\mbox{$^{\scriptscriptstyle(hh')}\!S$}} \def\mbox{$^{\scriptscriptstyle(h')}\!S$} {\mbox{$^{\scriptscriptstyle(h')}\!S$}} \def\mbox{\scriptsize(} {\mbox{\scriptsize(}} \def\mbox{$^{\scriptscriptstyle(\el h)}\!S$} {\mbox{$^{\scriptscriptstyle(\el h)}\!S$}} \def\mbox{sign} {\mbox{sign}} \newcommand\sm[2] {\mbox{\rm s}_{#1,#2}} \def\mbox{\rm s} {\mbox{\rm s}} \newcommand\Sm[2] {\mbox{\rm S}_{#1,#2}} \def\mbox{\rm S} {\mbox{\rm S}} \def$S$-matrix {$S$-matrix} \def\mbox{\scriptsize)} {\mbox{\scriptsize)}} \defsemi-simple {semi-simple} \def\scriptscriptstyle {\scriptscriptstyle} \def\scriptscriptstyle(\el\el'h) {\scriptscriptstyle(\ell\el'h)} \def\scriptscriptstyle(\el h) {\scriptscriptstyle(\ell h)} \def\scriptscriptstyle(h) {\scriptscriptstyle(h)} \def\scriptscriptstyle(hh') {\scriptscriptstyle(hh')} \def\scriptscriptstyle(h') {\scriptscriptstyle(h')} \def\sumihh {{\displaystyle\sum_{\scriptstyle i\atop \scriptstyle p_i|h',\,\scriptstyle p_i\not{\:|}\,h}}} \def\sumieh {{\displaystyle\sum_{\scriptstyle i\ne1\atop \scriptstyle p_i|h}}} \def\sumiezh {{\displaystyle\sum_{\scriptstyle i\ne1,2\atop \scriptstyle p_i|h}}} \newcommand\sumn[1] {{\displaystyle\sum_{#1=1}^{n}}} \newcommand\sumne[1] {{\displaystyle\sum_{#1=1}^{n-1}}} \newcommand\sumnee[1] {{\displaystyle\sum_{#1=1}^{n_1-1}}} \newcommand\sumneee[1]{{\displaystyle\sum_{#1=n_1}^{n_1'-1}}} \newcommand\sumnepe[1]{{\displaystyle\sum_{#1=1}^{n_1'-1}}} \newcommand\sumnie[1] {{\displaystyle\sum_{#1=1}^{n_i-1}}} \newcommand\sumnipe[1]{{\displaystyle\sum_{#1=1}^{n_i'-1}}} \def\sump {\sum_{\scriptstyle p\atop \scriptstyle n_p\ne0}} \newcommand\sumph[1] {{\displaystyle\sum_{#1\in{}^{\scriptscriptstyle(h)\!}P}}} \newcommand\sumpH[1] {\sum_{#1\in^{\scriptscriptstyle(h)\!}P}} \newcommand\sumphp[1] {{\displaystyle\sum_{#1\in{}^{\scriptscriptstyle(h')\!}P}}} \def\sumpl {\sum_{\scriptstyle p\atop \scriptstyle n'_p\ne0}} \newcommand\sumplh[1] {{\displaystyle\sum_{#1\in{}^{\scriptscriptstyle(\el h)\!}P}}} \newcommand\sumre[1] {{\displaystyle\sum_{#1=1}^r}} \newcommand\sumrE[1] {\sum_{#1=1}^r} \newcommand\sumro[1] {{\displaystyle\sum_{#1=0}^r}} \newcommand\sumrO[1] {\sum_{#1=0}^r} \renewcommand{\thesection.\arabic{equation}}{\thesection.\arabic{equation}} \def\mbox{$^{\sssh\!}T$} {\mbox{$^{\scriptscriptstyle(h)\!}T$}} \defthree-di\-men\-si\-o\-nal {three-di\-men\-si\-o\-nal} \def\tilde {\tilde} \def\topo {limit topology} \deftwo-di\-men\-si\-o\-nal {two-di\-men\-si\-o\-nal} \def{}^{\sss(\infty)\!}Q {{}^{\scriptscriptstyle(\infty)\!}Q} \defuntwisted affine Lie algebra {untwisted affine Lie algebra} \def{}^{\sss(\infty)\!}Q^- {{}^{\scriptscriptstyle(\infty)\!}Q^-} \newcommand\version[1] {\ifnum0=1 \typeout{}\typeout{#1}\typeout{} \vskip3mm \centerline{\fbox{\tt DRAFT -- #1 -- \today}} \vskip3mm \fi} \def\mbox{$ {\cal V}ir$} {\mbox{$ {\cal V}ir$}} \def{\rm vol} {{\rm vol}} \def\varphi^{} {\varphi^{}} \def\bar\varphi^{} {\bar\varphi^{}} \def{}^{\sssh\!}\varphi^{} {{}^{\scriptscriptstyle(h)\!}\varphi^{}} \def{}^{\sssh\!}\varphi {{}^{\scriptscriptstyle(h)\!}\varphi} \def{}^{\ssshp\!}\varphi^{} {{}^{\scriptscriptstyle(h')\!}\varphi^{}} \def{}^{\ssslh\!}\varphi^{} {{}^{\scriptscriptstyle(\el h)\!}\varphi^{}} \def{}^{\sss(\ho)\!}\varphi^{} {{}^{\scriptscriptstyle({h_\circ})\!}\varphi^{}} \def\mbox{$^{\sssh\!}W$} {\mbox{$^{\scriptscriptstyle(h)\!}W$}} \def\mbox{$^{\sss(h')\!}W$} {\mbox{$^{\scriptscriptstyle(h')\!}W$}} \def{}^{\sssh}W {{}^{\scriptscriptstyle(h)}W} \def\mbox{$^{\ssslh\!}W$} {\mbox{$^{\scriptscriptstyle(\el h)\!}W$}} \def\mbox{$^{\sssllh\!}W$} {\mbox{$^{\scriptscriptstyle(\el\el'h)\!}W$}} \def\mbox{$^{\sss(p)\!}W$} {\mbox{$^{\scriptscriptstyle(p)\!}W$}} \defwith respect to {with respect to } \defwith respect to the {with respect to the } \defWess\hy Zumino\hy Witten {Wess$\mbox{-\hspace{-.66 mm}-}$ Zumino$\mbox{-\hspace{-.66 mm}-}$ Witten} \defWZW model {WZW model} \defWZW theory {WZW theory} \defWZW theories {WZW theories} \def\hat\xi {\hat\xi} \def{\dl Z} {{\mathbb Z}} \def${\dl Z}$ {${\mathbb Z}$} \def\mbox{${\zet}_{\leq 0}$} {\mbox{${{\dl Z}}_{\leq 0}$}} \def\mbox{${\zet}_{>0}$} {\mbox{${{\dl Z}}_{>0}$}} \def\mbox{${\zet}_{\geq 0}$} {\mbox{${{\dl Z}}_{\geq 0}$}} \setlength{\unitlength}{.1 em} \def\begin{picture} {\begin{picture}} \def\begin{picture}(0,0) {\begin{picture}(0,0)} \def\end{picture} {\end{picture}} \newcommand\pictriangleupddr[6] {\begin{picture}(100,90) \put(46,70){#1} \put(-2,0){#2} \put(80,0){#3} \put(17,38){{#4}} \put(71,38){{#5}} \put(41,-4){{#6}} \put(47,65){\vector(-1,-2){25}} \put(53,65){\vector(1,-2){25}} \put(25,5){\vector(1,0){50}} \end{picture}} \newcommand\pictriangledowndd[3] {\begin{picture}(100,90) \put(44,0){#1} \put(17,38){{\footnotesize #2}} \put(72,38){{\footnotesize #3}} \put(47,15){\vector(-1,2){25}} \put(53,15){\vector(1,2){25}} \end{picture}} \def\number\month/\number\day/\number\year\ \ \ \hourmin {\number\month/\number\day/\number\year\ \ \ \hourmin } \def\pagestyle{draft}\thispagestyle{draft} \global\def0{1}{\pagestyle{draft}\thispagestyle{draft} \global\def0{1}} \global\def0{0} \catcode`\@=12 \documentclass[12pt]{article} \usepackage{amssymb,amsfonts} \def{\thesection.\arabic{equation}}{{\thesection.\arabic{equation}}} \setlength{\textwidth}{17cm} \setlength{\textheight}{23cm} \hoffset -17mm \topmargin= -18mm \raggedbottom \begin{document} \version\versionno \setlength{\unitlength}{.1 em} \begin{flushright} {\sf hep-th/9609124} \\ {\sf DESY 96-196} \\ {\sf IHES/P/96/47} \\ {\sf September 1996} \end{flushright} \begin{center} \vskip 1cm {\Large\bf WZW FUSION RINGS IN THE LIMIT} \vskip 2mm {\Large\bf OF INFINITE LEVEL} \vskip 12mm {{\large \ J\"urgen Fuchs} {\special{" 0 0 moveto 45 rotate}}}\\[5mm] {\small DESY}\\ {\small Notkestra\ss e 85}\\ {\small D -- 22603~~Hamburg} \\[5mm] and \\[5mm] {\large Christoph Schweigert}\\[5mm] {\small IHES}\\ {\small 35, Route de Chartres}\\ {\small F -- 91440 Bures-sur-Yvette} \end{center} \vskip 15mm \begin{quote} {\bf Abstract.} \\ We show that the WZW fusion rings at finite levels form a projective system with respect to the partial ordering provided by divisibility of the height, i.e.\ the level shifted by a constant. {}From this projective system we obtain WZW fusion rings in the limit of infinite level. This projective limit constitutes a mathematically well-defined prescription for the `classical limit' of WZW theories\ which replaces the naive idea of `sending the level to infinity'. The projective limit can be endowed with a natural topology, which plays an important r\^ole for studying its structure. The representation theory of the limit can be worked out by considering the associated fusion algebra; this way we obtain in particular an analogue of the Verlinde formula. \end{quote} \vfill{}{~}\\[1 mm]\noindent ------------------\\[1 mm]{} {\small {\special{" 0 0 moveto 45 rotate}}~~Heisenberg fellow} \newpage \Sect{Fusion rings}{wzw} Fusion rings constitute a mathematical structure which emerges in various contexts, for instance in the analysis of the superselection rules of two-di\-men\-si\-o\-nal\ quantum field theories; they describe in particular the basis independent contents of the operator product algebra\ of two-di\-men\-si\-o\-nal\ conformal field theories\ (for a review see \cite{jf24}). By definition, a {\em fusion ring\/} \mbox{${\cal R}$}\ is a unital commutative associative ring over the integers ${\dl Z}$\ which possesses the following properties: there is a distinguished basis $\mbox{${\cal B}$}=\{\varphi^{}_a\}$ which contains the unit and in which the structure constants \n abc\ are non-negative integers, and the evaluation at the unit induces an involutive automorphism, called the conjugation of \mbox{${\cal R}$}. A fusion ring is referred to as {\em rational\/} iff it is finite-dimensional. A rational fusion ring is called {\em modular\/} iff the matrix $S$ that diagonalizes simultaneously all fusion matrices \N a\ (i.e.\ the matrices with entries $(\N a)_b^{\ c} =\n abc$) is symmetric and together with an appropriate diagonal matrix $T$ generates a unitary representation\ of SL$(2,{\dl Z})$ (see e.g.\ \cite{kawA,jf24}). In this paper we consider the fusion rings of (chiral, unitary) WZW theories. A WZW theory\ is a conformal field theory\ whose chiral symmetry algebra\ is the semidirect sum of the Virasoro algebra\ with an untwisted affine Kac\hy Moo\-dy algebra\ \mbox{$\mathfrak g$}, with the level \mbox{$k_{}^{\scriptscriptstyle\vee}$}\ of the latter a fixed non-negative integer. To any untwisted affine Kac\hy Moo\-dy algebra\ \mbox{$\mathfrak g$}\ we can thus associate a family of fusion rings, parametrized by the level \mbox{$k_{}^{\scriptscriptstyle\vee}$}. The issue that we address in this paper is to construct an analogue of the WZW fusion ring for infinite level, which is achieved by giving a prescription for `sending the level to infinity' in an unambiguous manner. In view of the Lagrangian realization of WZW theories\ as sigma models, this procedure may be regarded as taking the `classical limit' of WZW theories. Performing a classical limit of a parametrized family of quantum field theories\ is a rather common concept in the path integral formulation of quantum theories; it simply corresponds to sending Planck's constant to zero, and hence provides a kind of correspondence principle. In the Lagrangian description of WZW theories\ as principal sigma models with Wess$\mbox{-\hspace{-.66 mm}-}$ Zumino terms, the r\^ole of Planck's constant is played by the inverse of the level \mbox{$k_{}^{\scriptscriptstyle\vee}$}\ of the underlying affine Lie algebra\ \mbox{$\mathfrak g$}. However, it is known that the path integral of a WZW sigma model strictly makes sense only if the level \mbox{$k_{}^{\scriptscriptstyle\vee}$}\ is an integer. In contrast to the path integral description, in the framework of algebraic approaches to quantum theory so far almost no attempts have been made to investigate limits of quantum field theories. In this paper we address this issue for the case of WZW theories. Now in an algebra ic treatment of WZW theories\ the integrality requirement just mentioned is immediately manifest. Namely, one observes that the structure of the theory depends sensitively on the value \mbox{$k_{}^{\scriptscriptstyle\vee}$}\ of the level. For non-negative integral \mbox{$k_{}^{\scriptscriptstyle\vee}$}\ the state space is a direct sum of unitary irreducible highest weight module s of the algebra \mbox{$\mathfrak g$}, but its structure changes quite irregularly when going from \mbox{$k_{}^{\scriptscriptstyle\vee}$}\ to $\mbox{$k_{}^{\scriptscriptstyle\vee}$}+1$; at intermediate, non-integral, values of the level there do not even exist any unitarizable highest weight representation s. These observations indicate that it is a rather delicate issue to define what is meant by the classical limit of a WZW theory, and it seems mandatory to perform this limit in a manner in which the level \mbox{$k_{}^{\scriptscriptstyle\vee}$}\ is manifestly kept integral (actually, treating the level formally as a continuous variable is potentially ambiguous even in situations where one deals with expressions which superficially make sense also at non-integral level). It must also be noted that a priori it is by no means clear whether the so obtained limit will be identical with or at least closely resemble the structures which originally served to define the quantum theory in terms of some classical field theory; in the case of WZW fusion rings, this underlying classical structure is the representation\ ring of the finite-dimensional\ simple Lie algebra\ $\mbox{$\bar{\mathfrak g}$}$ that is generated by the zero modes of the affine Lie algebra\ \mbox{$\mathfrak g$}. Indeed, it seems to be a quite generic feature of quantum theory that the classical limit does not simply reproduce the classical structure one started with. (Compare for instance the fact that in the path integral formulation of quantum field theory the classical paths are typically of measure zero in the space of all paths that contribute to the path integral. Similar phenomena also show up when the continuum limit of a lattice theory is constructed as a projective limit; see e.g.\ \cite{asle2,bellst}.) However, it is still reasonable to expect that the original classical structure is, in a suitable manner, contained in the classical limit; as we will see, this is indeed the case for our construction. The desire of being able to perform a limit in which the level tends to infinity stems in part from the fact that WZW theories\ and their fusion rings can be used to define a regularization of various systems (such as two-di\-men\-si\-o\-nal\ gauge theories or the Ponzano$\mbox{-\hspace{-.66 mm}-}$ Regge theory of simplicial three-di\-men\-si\-o\-nal\ gravity), with the unregularized system corresponding, loosely speaking, to the classical theory. As removing the regulator is always a subtle issue, it is mandatory that the limit of the regularized theory is performed in a well-defined, controllable manner, which, in addition, should preserve as much of the structure as possible. \medskip The basic idea which underlies our construction of the limit of WZW fusion rings is to interpret the collection of WZW fusion rings as a category ${\cal F}\!us\mbox{\scriptsize(}\mbox{$\mathfrak g$}\mbox{\scriptsize)}$\ within the category of all commutative rings and identify inside this category a projective system. By a standard category theoretic construction we can then obtain the limit (also known as the projective limit) of this projective system. The partial ordering underlying the projective system is based on a divisibility property of the parameter $\mbox{$k_{}^{\scriptscriptstyle\vee}$}+\mbox{$g_{}^{\scriptscriptstyle\vee}$}$ that together with the choice of the horizontal subalgebra\ \mbox{$\bar{\mathfrak g}$}\ characterizes the WZW theory\ (\mbox{$g_{}^{\scriptscriptstyle\vee}$}\ denotes the dual Coxeter number of \mbox{$\bar{\mathfrak g}$}; the sum $\mbox{$k_{}^{\scriptscriptstyle\vee}$}+\mbox{$g_{}^{\scriptscriptstyle\vee}$}$ is called the {\em height\/}). In contrast, in the literature often a purely formal prescription `\,$\mbox{$k_{}^{\scriptscriptstyle\vee}$}\to\infty$\,' is referred to as the classical limit of WZW theories. In that terminology it is implicit that the standard ordering on the set of levels is chosen to give it the structure of a directed set. Now the projective limit is associated to a projective system as a whole, not just to the collection of objects that appear in the system; in particular it depends on the underlying directed set and hence on the choice of partial ordering. Our considerations show, as a by-product, that it is not possible to associate to the standard ordering any well-defined limit of the fusion rings. The rest of this paper is organized as follows. We start in subsection \ref{swz} by introducing the category ${\cal F}\!us\mbox{\scriptsize(}\mbox{$\mathfrak g$}\mbox{\scriptsize)}$\ of WZW fusion rings associated to an untwisted affine Lie algebra\ \mbox{$\mathfrak g$}; in subsection \ref{s.qdim} conditions for the existence of non-trivial morphisms of this category are obtained. In subsection \ref{s.ps} we define the projective system, and in the remainder of section \ref{s.PS} we check that the morphisms introduced by this definition possess the required properties. The projective limit of the so obtained projective system is a unital commutative associative ring of countably infinite dimension. This ring \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}\ is constructed in \secref{pl}; there we also gather some basic properties of \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}\ and introduce a natural topology on \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}. In \secref{binf} a concrete description of a distinguished basis \mbox{$^{\scriptscriptstyle(\infty)\!\!}{\cal B}$}\ for the projective limit is obtained. This basis is similar to the distinguished bases of the fusion rings at finite level; in order to show that \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}\ is indeed generated by \mbox{$^{\scriptscriptstyle(\infty)\!\!}{\cal B}$}, the topology on \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}\ plays an essential r\^ole. In \secref{gb} we demonstrate that the representation\ ring of the horizontal subalgebra\ $\mbox{$\bar{\mathfrak g}$}\subset\mbox{$\mathfrak g$}$ is contained in the projective limit \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}\ as a proper subring. In the final \secref{rep} we study the representation\ theory of \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}, respectively\ of the associated fusion algebra over ${\mathbb C}$\,. In particular, we determine all irreducible representation s and show that \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}\ possesses a property which is the topological analogue of semi-simplicity, namely that any continuous \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}-module is the closure of a direct sum of irreducible modules. To obtain these results it is again crucial to treat the projective limit as a topological space. Finally, we establish an analogue of the Verlinde formula for \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}. \Sect{The projective system of WZW fusion rings}{PS} \subsection{WZW fusion rings}\label{swz} The primary fields of a unitary WZW theory\ are labelled by integrable highest weights of the relevant affine Lie algebra\ \mbox{$\mathfrak g$}, or what is the same, by the value \mbox{$k_{}^{\scriptscriptstyle\vee}$}\ of the level and by dominant integral weights $\Lambda$ of \mbox{$\bar{\mathfrak g}$}\ (the horizontal subalgebra of \mbox{$\mathfrak g$}) whose inner product with the highest coroot of \mbox{$\bar{\mathfrak g}$}\ is not larger than \mbox{$k_{}^{\scriptscriptstyle\vee}$}. We denote by \mbox{$g_{}^{\scriptscriptstyle\vee}$}\ the dual Coxeter number of the simple Lie algebra\ \mbox{$\bar{\mathfrak g}$}\ and define \begin{equation} \mbox{$I$}:= \{ i\in{\dl Z} \mid i\ge\mbox{$g_{}^{\scriptscriptstyle\vee}$} \} \,. \labl i Thus \mbox{$I$}\ is the set of possible values of the {\em height\/} $h\equiv \mbox{$k_{}^{\scriptscriptstyle\vee}$}+\mbox{$g_{}^{\scriptscriptstyle\vee}$}$ of the WZW theory\ based on \mbox{$\mathfrak g$}. For any $h\iN\II$ the fusion rules of a WZW theory\ at height $h$ define a modular fusion ring, with the elements of the distinguished basis corresponding to the primary fields. We denote this ring by \mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$}\ and its distinguished basis by \mbox{$^{\scriptscriptstyle(h)}\!{\cal B}$}, and the corresponding generators of SL$(2,{\dl Z})$ by \mbox{$^{\scriptscriptstyle(h)}\!S$}\ and \mbox{$^{\sssh\!}T$}. The distinguished basis \mbox{$^{\scriptscriptstyle(h)}\!{\cal B}$}\ of the ring \mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$}\ can be labelled as \begin{equation} \mbox{$^{\scriptscriptstyle(h)}\!{\cal B}$}=\{{}^{\sssh\!}\varphi^{}_a \,|\,a\!\in\!\mbox{$_{}^{\scriptscriptstyle(h)\!\!}P$}\} \end{equation} by the set \begin{equation} \mbox{$_{}^{\scriptscriptstyle(h)\!\!}P$} := \{ a\in\mbox{$\overline L^{\rm w}$} \mid a^i\ge1\ {\rm for}\ i=1,2,...\,,r;\; (a,\theta^{\scriptscriptstyle\vee})<h \} \labl{ph} of integral weights in the interior of (the horizontal projection of) the fundamental Weyl chamber of \mbox{$\mathfrak g$}\ at {\em level\/} $h$; here $r$, $\theta^{\scriptscriptstyle\vee}$ and \mbox{$\overline L^{\rm w}$}\ denote the rank, the highest coroot and the weight lattice of \mbox{$\bar{\mathfrak g}$}, respectively. Note that from here on we use shifted \mbox{$\bar{\mathfrak g}$}-weights $a=\Lambda+\rho$, which have level $h =\mbox{$k_{}^{\scriptscriptstyle\vee}$}+\mbox{$g_{}^{\scriptscriptstyle\vee}$}$, in place of unshifted weights $\Lambda$ which are at level \mbox{$k_{}^{\scriptscriptstyle\vee}$}. Here $\rho$ is the Weyl vector of \mbox{$\bar{\mathfrak g}$}; in particular, $a=\rho$ is the label of the unit element of \mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$}. This convention will simplify various formul\ae\ further on. The ring product of \mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$}\ will be denoted by the symbol `\,$\star$\,'; thus the fusion rules are written as \begin{equation} {}^{\sssh\!}\varphi^{}_a \star {}^{\sssh\!}\varphi^{}_b = \sumph c \nh abc\,{}^{\sssh\!}\varphi^{}_c \,. \end{equation} The collection $(\mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$})_{h\in\II}$ of such WZW fusion rings forms a category, more precisely a subcategory of the category of commutative rings, which we denote by ${\cal F}\!us\mbox{\scriptsize(}\mbox{$\mathfrak g$}\mbox{\scriptsize)}$. The objects of ${\cal F}\!us\mbox{\scriptsize(}\mbox{$\mathfrak g$}\mbox{\scriptsize)}$\ are the rings \mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$}, and the morphisms (arrows) are those ring homomorphisms (which are automatically unital and compatible with the conjugation) which map the basis \mbox{$^{\scriptscriptstyle(h')}\!{\cal B}$}\ up to sign factors to \mbox{$^{\scriptscriptstyle(h)}\!{\cal B}$}. These are the natural requirements to be imposed on morphisms. Namely, one preserves precisely all structural properties of the fusion ring, except for the positivity of the structure constants; the latter is not an algebraic property, so that one should be prepared to give it up. \subsection{Existence of morphisms}\label{s.qdim} It is not a priori clear whether the category ${\cal F}\!us\mbox{\scriptsize(}\mbox{$\mathfrak g$}\mbox{\scriptsize)}$\ as defined above has any non-trivial morphisms at all. To analyze this issue, we consider the quotients ${\mbox{$^{\scriptscriptstyle(h)}\!S$}_{a,b}} /{\mbox{$^{\scriptscriptstyle(h)}\!S$}_{a,\rho}}$ ($a,b\!\in\!\mbox{$_{}^{\scriptscriptstyle(h)\!\!}P$}$) of $S$-matrix\ elements. These are known as the (generalized) quantum dimension s, or more precisely, as the $a$th quantum dimension\ of the element ${}^{\sssh\!}\varphi^{}_b$, of the modular fusion ring \mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$}. The generalized quantum dimension s furnish precisely all inequivalent irreducible representation s of \mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$}\ \cite{kawA}. We denote by \begin{equation} \qd ha :\quad \mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$}\to{\mathbb C}\,, \quad {}^{\sssh\!}\varphi^{}_b\mapsto \frac{\mbox{$^{\scriptscriptstyle(h)}\!S$}_{a,b}}{\mbox{$^{\scriptscriptstyle(h)}\!S$}_{a,\rho}} \labl{qd} the irreducible representation\ of \mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$}\ which associates to any element its $a$th generalized quantum dimension. Assume now that $f\!:\;\mbox{$^{\scriptscriptstyle(h')\!}{\cal R}$}\to\mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$}$ is a non-trivial morphism, i.e.\ a ring homomorphism which maps the distinguished basis \mbox{$^{\scriptscriptstyle(h')}\!{\cal B}$}\ of \mbox{$^{\scriptscriptstyle(h')\!}{\cal R}$}\ to the basis \mbox{$^{\scriptscriptstyle(h)}\!{\cal B}$}\ of \mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$}. Then the composition $\qd ha\circ f$ provides us with a one-dimensional, and hence irreducible, representation\ of \mbox{$^{\scriptscriptstyle(h')\!}{\cal R}$}, i.e.\ we have $\qd ha\circ f=\qd {h'}{a'}$ for some $a'\!\in\!\mbox{$_{}^{\scriptscriptstyle(h')\!\!}P$}$. Let now \mbox{$^{\scriptscriptstyle(h)\!\!}L$}\ denote the extension of the field ${\mathbb Q}$ of rational numbers by the quantum dimension s\ $\mbox{$^{\scriptscriptstyle(h)}\!S$}_{a,b}/\mbox{$^{\scriptscriptstyle(h)}\!S$}_{a,\rho}$ of all elements of \mbox{$^{\scriptscriptstyle(h)}\!{\cal B}$}; the observation just made then implies that \begin{equation} \mbox{$^{\scriptscriptstyle(h)\!\!}L$} \subseteq \mbox{$^{\scriptscriptstyle(h')}\!L$} \labl{ll} (when $f$ is surjective, one gets in fact the whole field $\mbox{$^{\scriptscriptstyle(h)\!\!}L$}$). As we will see, this result puts severe constraints on the existence of morphisms from \mbox{$^{\scriptscriptstyle(h')\!}{\cal R}$}\ to \mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$}. It follows from the Kac\hy Peterson formula\ \cite{KAc3} for the $S$-matrix\ that \begin{equation} \mbox{$^{\scriptscriptstyle(h)\!\!}L$}\subseteq\qzm h\,, \labl q with $\zeta_m:=\exp(2\pi{\rm i}/m)$ and $M$ the smallest positive integer for which all entries of the metric on the weight space of \mbox{$\bar{\mathfrak g}$}\ are integral multiples of $1/M$ (except for $\mbox{$\bar{\mathfrak g}$}=A_r$ where $M=r+1$, $M$ satisfies $M\leq4$). The inclusion \erf{ll} therefore implies that \mbox{$^{\scriptscriptstyle(h)\!\!}L$}\ lies in the intersection $\qzm h \cap \qzm{h'} = \qzm{\,\ggt h{h'}}$, and that this intersection is strictly larger than ${\mathbb Q}$ unless $\mbox{$^{\scriptscriptstyle(h)\!\!}L$}={\mathbb Q}$. Here $\ggt mn$ stands for the largest common divisor of $m$ and $n$. In the specific case that $h$ and $h'$ are coprime, $\ggt h{h'}=1$, it follows that \begin{equation} \mbox{$^{\scriptscriptstyle(h)\!\!}L$} \,\subseteq\, \mbox{$^{\scriptscriptstyle(h')}\!L$} \cap \qzm h \,\subseteq\, \qzm{} \,. \end{equation} Now typically the field \mbox{$^{\scriptscriptstyle(h)\!\!}L$}\ is quite a bit smaller than \qzm h, i.e.\ the inequality \erf q is not saturated (e.g.\ if the ring is self-conjugate, \mbox{$^{\scriptscriptstyle(h)\!\!}L$}\ is already contained in the maximal real subfield of \qzm h); nevertheless, inspection shows that the requirement $\mbox{$^{\scriptscriptstyle(h)\!\!}L$}\subseteq\qzm{}$ is fulfilled only in very few cases (for instance, for \mbox{$\bar{\mathfrak g}$}\ of type $B_{2n},\,C_r,\,D_{2n},\, E_7,\,E_8$ or $F_4$, one has $M\le2$ so that the requirement is just $\mbox{$^{\scriptscriptstyle(h)\!\!}L$}= {\mathbb Q}$). In addition, the main quantum dimension s $\mbox{$^{\scriptscriptstyle(h)}\!S$}_{a,\rho}/\mbox{$^{\scriptscriptstyle(h)}\!S$}_{\rho,\rho}$ lie in fact in \qz{2h}, and hence the above requirement would restrict them to lie in $\qz{2h}\cap\qzm{}=\qz{\ggt{2h}M}$, and thus to be rational whenever $2h$ and $M$ are coprime. It follows that for almost all pairs $h,\,h'$ of coprime heights there cannot exist any morphism from \mbox{$^{\scriptscriptstyle(h')\!}{\cal R}$}\ to \mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$}. The same arguments also show that the existence of non-trivial morphisms becomes the more probable the larger the value of $\ggt h{h'}$ is. The most favourable situation is when $h'$ is a multiple of $h$; in the next section we will show that in this case a whole family of morphisms from \mbox{$^{\scriptscriptstyle(h')\!}{\cal R}$}\ to \mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$}\ (with $h\iN\II$ arbitrary) can be constructed in a natural way. The considerations above indicate in particular that the naive way of taking the limit `\,$k\to\infty$\,' with the standard ordering on the set \mbox{$I$}\ cannot correspond to any well-defined limit of the WZW fusion rings. In contrast, as we will show, when replacing the standard ordering by a suitable partial ordering, a limit can indeed be constructed, namely as the projective limit of a projective system that is associated to that partial ordering. Let us also mention that the required ring homomorphism property implies that any morphism of ${\cal F}\!us\mbox{\scriptsize(}\mbox{$\mathfrak g$}\mbox{\scriptsize)}$\ maps simple currents to simple currents. (By definition, simple currents are those elements $\varphi_a$ of the distinguished basis \mbox{$^{\scriptscriptstyle(h)}\!{\cal B}$}\ which have inverses in the fusion ring; they satisfy $\sum_c\n abc=1$ for all $b$. Such elements are sometimes also called units of the ring, not to be confused with the unit element of the fusion ring.) \subsection{The projective system}\label{s.ps} On the set \mbox{$I$}\ \erf i of heights one can define a partial ordering `\,$\preceq$\,' by \begin{equation} i \preceq j \ :\Leftrightarrow\ i\,|\,j \,, \labl l where the vertical bar stands for divisibility. For any two elements $i,i'\iN\II$ there then exists a $j\iN\II$ (for example, the smallest common multiple of $i$ and $i'$) such that $i\preceq j$ and $i'\preceq j$. Thus the partial ordering \erf l endows \mbox{$I$}\ with the structure of a {\em directed set}. We will now show that when the set \mbox{$I$}\ is considered as a directed set via the partial ordering \erf l, the collection $(\mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$})_{h\in\II}$ of WZW fusion rings can be made into a {\em projective system}, that is, for each pair $i,j\iN\II$ satisfying $i\preceq j$ there exists a morphism \begin{equation} \ff ji:\quad \mbox{$^{\scriptscriptstyle(j)\!}{\cal R}$}\to\mbox{$^{\scriptscriptstyle(i)\!}{\cal R}$}\,, \labl f such that \ff ii is the identity for all $i\!\in\! I$ and such that for all $i,j,k\iN\II$ which satisfy $i\preceq j\preceq k$, the diagram \begin{equation} \begin{array}{l} {}\\[-2.2em] \pictriangleupddr {$\!\!\!$\mbox{$^{\scriptscriptstyle(k)\!}{\cal R}$}}{\mbox{$^{\scriptscriptstyle(j)\!}{\cal R}$}~~}{\mbox{$^{\scriptscriptstyle(i)\!}{\cal R}$}}{\fF kj}{\fF ki}{\fF ji} \\[.3em] \end{array} \labl1 commutes. We have to construct the maps \ff ij for all pairs $i,\,j$ with $i|j$. Writing $i=h$ and $j=\ell h$ with $\ell\!\in\!{\dl N}$, the construction goes as follows. The horizontal projection \mbox{$^{\sssh\!}W$}\ of the affine Weyl group at height $h$ has the structure of a semidirect product $\mbox{$^{\sssh\!}W$} = \overline W\semitimes h\mbox{$\overline L^{\scriptscriptstyle\vee}$}$, with $\overline W$ the Weyl group and \mbox{$\overline L^{\scriptscriptstyle\vee}$}\ the coroot lattice of \mbox{$\bar{\mathfrak g}$}, so that in particular \mbox{$^{\sssh\!}W$}\ is contained as a finite index subgroup in \mbox{$^{\ssslh\!}W$}, the index having the value $\ell^{\,r}_{}$. Thus any orbit of \mbox{$^{\sssh\!}W$}\ decomposes into orbits of \mbox{$^{\ssslh\!}W$}, and each Weyl chamber at height $\ell h$ is the union of $\ell^{\,r}_{}$ Weyl chambers at height $h$. As a consequence, we find that the following statement holds for the set \mbox{$_{}^{\scriptscriptstyle(\el h)\!\!}P$}\ defined according to \erf{ph}. To any $a\!\in\!\mbox{$_{}^{\scriptscriptstyle(\el h)\!\!}P$}$ there either exists a unique element $w_a\!\in\!\mbox{$^{\sssh\!}W$}$ such that \begin{equation} a':=w_a(a) \labl{a'} belongs to the set \mbox{$_{}^{\scriptscriptstyle(h)\!\!}P$}, or else $a$ lies on the boundary of some affine Weyl chamber at height $h$. In the former case we define \begin{equation} \ff{\el h}h({}^{\ssslh\!}\varphi^{}_a):= \epsilon_{\el}(a)\cdot {}^{\sssh\!}\varphi^{}_{a'} \labl' with $\epsilon_{\el}(a)=\mbox{sign}(w_a),$ while in the latter case we set $\ff{\el h}h({}^{\ssslh\!}\varphi^{}_a):=0$. It is convenient to include this latter case into the formula \erf', which is achieved by setting \begin{equation} \epsilon_{\el}(a):= \left\{ \begin{array}{ll} 0 & \mbox{if $a$ lies on the boundary of an}\\ & \mbox{affine Weyl chamber at height $h$\,,} \\[1mm] \mbox{sign}(w_a) & {\rm else} \,. \end{array}\right. \labl{eps} \subsection{Proof of the morphism properties} We have to prove that $\ff ij$ defined this way is a ring homomorphism and that it satisfies the composition property \erf1. It is obvious from the definition that $\ff ii={\sf id}$ (and also that $\ff ij$ is surjective). To show the homomorphism property, we write $\ff ij$ in matrix notation, and for convenience use capital letters for the fusion ring \mbox{$^{\scriptscriptstyle(\el h)\!}{\cal R}$}\ and lower case letters for the fusion ring \mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$}. Thus the elements of the basis \mbox{$^{\scriptscriptstyle(\el h)}\!{\cal B}$}\ of \mbox{$^{\scriptscriptstyle(\el h)\!}{\cal R}$}\ are denoted by $\phi^{}_A\equiv{}^{\ssslh\!}\varphi^{}_{\!A}$ with $A\!\in\!\mbox{$_{}^{\scriptscriptstyle(\el h)\!\!}P$}$, while for the elements of \mbox{$^{\scriptscriptstyle(h)}\!{\cal B}$}\ we just write $\varphi^{}_a$ with $a\!\in\!\mbox{$_{}^{\scriptscriptstyle(h)\!\!}P$}$, and we use the notation \mbox{\rm S}\ and \mbox{\rm s}\ for the $S$-matrices in place of \mbox{$^{\scriptscriptstyle(\el h)}\!S$}\ and \mbox{$^{\scriptscriptstyle(h)}\!S$}, respectively. The mapping is then defined on the preferred basis $\mbox{$^{\scriptscriptstyle(\el h)}\!{\cal B}$}$ as \begin{equation} \ff{\el h}h(\phi^{}_A) = \sumph b \P Ab\, \varphi^{}_b \, \labl3 with \begin{equation} \P Ab \equiv {}^{\scriptscriptstyle(\ell h,h)\!\!}_{}\P Ab := \epsilon_{\el}(A)\, \delta^{}_{w_A(A),b} \,, \labl P and extended linearly to all of \mbox{$^{\scriptscriptstyle(\el h)\!}{\cal R}$}. As has been established in \cite{fusS2}, the matrix \erf P satisfies the relations \futnote{In \cite{fusS2}, mappings of the type \erf' were encountered as so-called quasi-Galois scalings. In that setting, the level of the WZW theory\ is not changed, while the weights $A$ are scaled by a factor of $\ell$, followed by an appropriate affine Weyl transformation to bring the weight $\ell A$ back to the Weyl chamber \mbox{$_{}^{\scriptscriptstyle(\el h)\!\!}P$}\ or to its boundary. Since what matters is only the relative `size' of weights and the translation part of the Weyl group, these mappings are effectively the same as in the present setting where there is no scaling of the weights but the extension from \mbox{$^{\ssslh\!}W$}\ to \mbox{$^{\sssh\!}W$}\ scales the translation lattice down by a factor of $\ell$. \\[.2em]Note that in \cite{fusS2} the letter $P$ was used for the matrix \erf P in place of $F$, and $D$ was defined as the transpose of the matrix \erf D.} \begin{equation} \mbox{\rm S}\,F = \ell^{\,r/2}_{}\,D\,\mbox{\rm s} \,, \qquad F\,\mbox{\rm s} = \ell^{\,r/2}_{}\,\mbox{\rm S}\,D \,, \labl{sS} with \begin{equation} \D Ab \equiv {}^{\scriptscriptstyle(\ell h,h)\!\!}_{}\D Ab:= \delta_{A,\ell b}\,. \labl D Furthermore, from the Kac$\mbox{-\hspace{-.66 mm}-}$ Peterson formula \cite{KAc3} for the modular matrix $S$, one deduces the identity \begin{equation} \sm ab = \ell^{\,r/2}_{}\, \Sm{\ell a}b \labl K for all $a,b\!\in\!\mbox{$_{}^{\scriptscriptstyle(h)\!\!}P$}$. Combining the relations \erf P -- \erf K and the Verlinde formula \cite{verl2}, we obtain for any pair $A,B\!\in\!\mbox{$_{}^{\scriptscriptstyle(\el h)\!\!}P$}$ \begin{equation} \begin{array}{l} \!\! \ff{\el h}h(\phi^{}_A \star \phi^{}_B) = \sumplh C \nlh ABC\, \ff{\el h}h(\phi^{}_C) = \sumplh{C,D}\;\sumph e \displaystyle\frac {\Sm AD^{} \Sm BD^{} \Sm CD^*}{\Sm\rho D} \, \P Ce\, \varphi^{}_e \\{}\\[-2.6mm] \hsp{23.2} = \!\sumplh{D}\;\sumph e \displaystyle\frac {\Sm AD \Sm BD (\mbox{\rm S}^*\!F)_{D,e}}{\Sm\rho D} \, \varphi^{}_e = \ell^{\,r/2}_{}\cdot \!\!\!\sumplh{D}\;\sumph e \displaystyle\frac {\Sm AD^{} \Sm BD^{} (D\mbox{\rm s}^*)_{D,e}}{\Sm\rho D} \, \varphi^{}_e \\{}\\[-2.6mm] \hsp{23.2} = \ell^{\,r/2}_{}\cdot \sumph{d,e} \displaystyle\frac {\Sm A{\ell d}^{} \Sm B{\ell d}^{} \,\sm de^*} {\Sm \rho{\ell d}}\, \varphi^{}_e = \ell^{\,r}_{}\cdot \sumph{d,e} \displaystyle\frac {\Sm A{\ell d}^{} \Sm B{\ell d}^{} \,\sm de^*} {\sm \rho d}\, \varphi^{}_e \\{}\\[-2.6mm] \hsp{23.2} = \ell^{\,r}_{}\cdot \sumph{d,e} \displaystyle\frac {(\SMD)_{A,d}^{} (\SMD)_{B,d}^{} \, \sm de^*} {\sm \rho d}\, \varphi^{}_e = \sumph{d,e} \displaystyle\frac {(F\mbox{\rm s})_{A,d}^{} (F\mbox{\rm s})_{B,d}^{} \,\sm de^*} {\sm \rho d}\, \varphi^{}_e \\{}\\[-2.6mm] \hsp{23.2} = \sumph{a,b,d,e} \P Aa \P Bb\, \displaystyle\frac {\sm ad^{} \,\sm bd^{} \,\sm ed^*} {\sm \rho d}\, \varphi^{}_e = \sumph{a,b,c} \P Aa \P Bb\, \nh abc \, \varphi^{}_c \\{}\\[-2.2mm] \hsp{23.2} = \sumph{a,b} \P Aa \P Bb\, \varphi^{}_a \star \varphi^{}_b = \ff{\el h}h(\phi_A) \star \ff{\el h}h(\phi_B) \end{array} \end{equation} Thus $\ff{\el h}h$ is indeed a homomorphism. As a side remark, let us mention that an analogous situation arises for the conformal field theories\ which describe a free boson compactified on a circle of rational radius squared. These theories are labelled by an (even) positive integer $h$, and for each value of $h$ the fusion ring is just the group ring ${\dl Z}\zet_h$ of the abelian group ${\dl Z}_h={\dl Z}/h{\dl Z}$. The modular $S$-matrix\ is given by $\mbox{$^{\scriptscriptstyle(h)}\!S$}_{p,q} = h^{-1/2} \exp(2\pi{\rm i} p q /h)$, where the labels $p$ and $q$ which correspond to the primary fields are integers which are conveniently considered as defined modulo $h$. It is straightforward to check that the identities \erf{sS} are again valid (with $r$ set to 1) if one defines ${}^{\scriptscriptstyle(\ell h,h)\!}_{}\P Ab:=\delta^{(h)}_{A,b}$ and ${}^{\scriptscriptstyle(\ell h,h)\!}_{}\D Ab:=\delta^{(\ell h)}_{A,\ell b}$, where the superscript on the $\delta$-symbol $\delta^{(m)}_{a,b}$ indicates that equality needs to hold only modulo $m$. As a consequence, this way we obtain again a projective system based on the divisibility of $h$ (the composition property is immediate). Moreover, precisely as in the case of WZW theories, with a different partial ordering of the set $\{h\}=\mbox{${\zet}_{>0}$}$ it is not possible to define a projective system. \subsection{Proof of the composition property} Finally, the composition property \erf1 of the homomorphisms \erf' is equivalent to the relation \begin{equation} \sumplh B {}^{\scriptscriptstyle(\ell\el'h,\ell h)\!}_{}\P {\sf A}B \, {}^{\scriptscriptstyle(\ell h,h)\!}_{}\P Bc = {}^{\scriptscriptstyle(\ell\el'h,h)\!}_{}\P {\sf A}c \labl{ppp} among the projection matrices $F$ that involve the three different heights $h$, $\ell h$ and $\ell\el'h$. Here as before the elements of \mbox{$_{}^{\scriptscriptstyle(h)\!\!}P$}\ and \mbox{$_{}^{\scriptscriptstyle(\el h)\!\!}P$}\ are denoted by lower case and capital letters, respectively, while for the elements of \mbox{$_{}^{\scriptscriptstyle(\el\el'h)\!\!}P$}\ we use sans-serif font. The relation \erf{ppp} is in fact an immediate consequence of the definition of the homomorphisms \ff ij. The explicit proof is not very illuminating; the reader who wishes to skip it should proceed directly to \secref{pl}. To prove \erf{ppp}, let us first assume that the left hand side does not vanish. Then there exist unique Weyl transformations $\overline w_1,\overline w_2 \!\in\!\overline W$ and unique vectors $\beta_1,\beta_2\!\in\! h\mbox{$\overline L^{\scriptscriptstyle\vee}$}$ in the coroot lattice scaled by $h$, and a unique weight $B\!\in\!\mbox{$_{}^{\scriptscriptstyle(\el h)\!\!}P$}$, such that \begin{equation} \overline w_1({\sf A}) + \ell\beta_1 = B \,, \qquad \overline w_2(B) +\beta_2 = c \,, \labl{ww} and the left hand side of \erf{ppp} takes the value \begin{equation} \epsilon_{\el\el'}({\sf A})\,\epsilon_{\el}(B) = \mbox{sign}(\overline w_1)\,\mbox{sign}(\overline w_2) = \mbox{sign}(\overline w_1\overline w_2) \,. \labl{sww} By combining the two relations \erf{ww}, it follows that \begin{equation} \overline w_2\overline w_1({\sf A}) + \beta = c \,, \labl{beta} where $\beta=\ell\,\overline w_2(\beta_1)+\beta_2$. Since $\beta$ is again an element of $h\mbox{$\overline L^{\scriptscriptstyle\vee}$}$, this means that \erf{beta} describes, up to sign, the mapping corresponding to the right hand side of \erf{ppp}. Further, the sign of the right hand side is then given by $\mbox{sign}(\overline w_2\overline w_1)$ and hence equal to \erf{sww}; thus \erf{ppp} indeed holds. We still have to analyze \erf{ppp} when its left hand side vanishes. Then either the $\sf A$\,th row of ${}^{\scriptscriptstyle(\ell\el'h,\ell h)\!}_{}F$ or the $c$\,th column of ${}^{\scriptscriptstyle(\ell h,h)\!}_{}F$ must be zero. In the former case, the weight ${\sf A}\!\in\!\mbox{$_{}^{\scriptscriptstyle(\el\el'h)\!\!}P$}$ belongs to the boundary of some Weyl chamber with respect to \mbox{$^{\ssslh\!}W$}, and thus there exist $\overline w\!\in\!\overline W$ and $\beta\!\in\! h\mbox{$\overline L^{\scriptscriptstyle\vee}$}$ such that $\overline w({\sf A})+\ell\beta={\sf A}$. But this means that ${\sf A}\!\in\!\mbox{$_{}^{\scriptscriptstyle(\el\el'h)\!\!}P$}$ also lies on the boundary of some Weyl chamber with respect to $\mbox{$^{\sssh\!}W$}\supset\mbox{$^{\ssslh\!}W$}$, and hence also the right hand side of \erf{ppp} vanishes as required. In the second case, there are unique elements $w_1\!\in\!\mbox{$^{\ssslh\!}W$}$ and $w_2\!\in\!\mbox{$^{\sssh\!}W$}$ satisfying $w_1({\sf A})=B$ and $w_2(B)=B$. Because of $\mbox{$^{\ssslh\!}W$}\subset\mbox{$^{\sssh\!}W$}$, $w_1$ can also be considered as an element of the Weyl group \mbox{$^{\sssh\!}W$}\ at height $h$. By assumption, $w_2$ is a non-trivial element of \mbox{$^{\sssh\!}W$}. The product $w_0:=w_1^{-1}w_2^{}w_1^{}\!\in\!\mbox{$^{\sssh\!}W$}$ is then non-trivial, too, and satisfies \begin{equation} w_0({\sf A}) =w_1^{-1}w_2^{}w_1^{}({\sf A}) =w_1^{-1}w_2^{}(B)=w_1^{-1}(B)={\sf A} \,. \end{equation} Thus the weight $\sf A$ is invariant under a non-trivial element of \mbox{$^{\sssh\!}W$}\ and hence lies on the boundary of some Weyl chamber with respect to \mbox{$^{\sssh\!}W$}; this implies again that the right hand side of \erf{ppp} vanishes as required. This concludes the proof of \erf{ppp}, and hence of the claim that $(\mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$})_{h\in\II}$ together with the maps \ff ij constitutes a projective system. \Sect{The projective limit \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}}{pl} We are now in a position to construct the projective limit \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}\ of the projective system that we introduced in subsection \ref{s.ps}. \subsection{Projective limits and coherent sequences} A projective system $(\mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$})_{h\in I}$\, in some category \mbox{${\cal C}$}\ is said to possess a {\em projective limit\/} $(\mbox{$\cal L$},f)$ (also called the inverse limit, or simply the limit) if there exist an object \mbox{$\cal L$}\ as well as a family $f$ of morphisms $\ef h\!:\ \mbox{$\cal L$}\to\mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$}$ (for all $h\!\in\!\mbox{$I$}$) which satisfy the following requirements (see e.g.\ \cite{ENCY}). First, for all $h,h'\!\in\!\mbox{$I$}$ with $h\preceq h'$ the diagram \begin{equation} \begin{array}{l} {}\\[-2.2em] \pictriangleupddr {\mbox{$\cal L$}}{\mbox{$^{\scriptscriptstyle(h')\!}{\cal R}$}~~}{\mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$}}{\eF {h'}}{\eF h}{\fF{h'}h} \\[.3em] \end{array} \labl I commutes; and second, the following {\em universal property\/} holds: for any object \mbox{$\cal O$}\ of the category for which a family of morphisms $\,\eg h\!:\;\mbox{$\cal O$}\to\mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$}\,$ ($h\!\in\!\mbox{$I$}$) exists which possesses a property analogous to \erf I, i.e.\ \begin{equation} \ff {h'}h \circ \eg{h'} = \eg h \quad {\rm for}\ \, h\preceq h' \, , \end{equation} there exists a unique morphism $\,g\!:\;\mbox{$\cal O$}\to\mbox{$\cal L$}$ such that the diagram \begin{equation} \begin{array}{l} \begin{picture}(120,170) \put(0,90){ \pictriangleupddr {\mbox{$\cal L$}}{\mbox{$^{\scriptscriptstyle(h')\!}{\cal R}$}~~}{\mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$}}{\eF{h'}}{\eF h}{\fF{h'}h} } \put(0,18){ \pictriangledowndd {\,\mbox{$\cal O$}}{\eG{h'}}{\eG h} } \put(113,92){\begin{picture}(0,0) \put(0,0){\oval(30,142)[r]} \put(19.8,3.5){$\scsg$} \put(0,71){\vector(-1,0){45}} \put(0,-71){\line(-1,0){45}} \end{picture}} \end{picture} \\[-1.8em] \end{array} \labl{II} commutes for all $h,h'\!\in\!\mbox{$I$}$ with $h\preceq h'$. To be precise, in the above characterization of the projective limit $(\mbox{$\cal L$},f)$ it is implicitly assumed that \mbox{$\cal L$}\ is an object in \mbox{${\cal C}$}\ and that the \ef i are morphisms of \mbox{${\cal C}$}. But in fact such an object and such morphisms need not exist. In the general case one must rather employ a definition of the projective limit as a certain functor from the category \mbox{${\cal C}$}\ to the category of sets, and then the question arises whether this functor is `representable' through an object \mbox{$\cal L$}\ and morphisms \ef i as described above. In this language the crucial issue is the existence of a representing object \mbox{$\cal L$}\ (see e.g.\ \cite{ARti,PAre,HIst}). \futnot{PAre': sec. 2.5; HIst: chap. VIII.5} Now one and the same projective system can frequently be regarded as part of various different categories. For instance, when describing the projective system of our interest one can restrict oneself to the category ${\cal F}\!us\mbox{\scriptsize(}\mbox{$\mathfrak g$}\mbox{\scriptsize)}$. As we will see, when doing so a projective limit of the projective system does not exist. But one can also consider it, say, in the category of commutative rings, or in the still bigger category of vector spaces, or even in the category of sets. The existence and the precise form of the projective limit usually depend on the choice of category. In our case, however, the category \mbox{${\cal C}$} \,=\,${\cal F}\!us\mbox{\scriptsize(}\mbox{$\mathfrak g$}\mbox{\scriptsize)}$\ we start with is small, i.e.\ its objects are sets, and as a consequence there exists a natural construction by which the object \mbox{$\cal L$}\ and the morphisms \ef i can be obtained in a concrete manner (in particular, \mbox{$\cal L$}\ is again a set). Moreover, it turns out that the projective limit we obtain in the category of sets is exactly the same as the limit that we obtain in the category of commutative rings or vector spaces, which also indicates that this way of performing the limit is a quite natural. This construction proceeds as follows. Given a projective system of objects \mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$}\ and morphisms \ff{h'}h of a small category \mbox{${\cal C}$}, one regards the objects $\mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$}\!\in\!\mbox{${\cal C}$}$ as sets and considers the infinite direct product $\prod_{h\inI}\!\mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$}$ of all objects of \mbox{${\cal C}$}. The elements of this set are those maps \begin{equation} \psi:\ \; \mbox{$I$} \;\to\; \bigcup_{h\inI}^. \mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$} \labl{psi} from the index set \mbox{$I$}\ to the disjoint union of all objects \mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$}\ which obey $\psi(h)\!\in\!\mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$}$ for all $h\!\in\!\mbox{$I$}$; they are sometimes called `generalized sequences' (ordinary sequences can be formulated in this language by considering the index set ${\mathbb N}$ with the standard ordering $\le$\,). The subset $\mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$} \subset\prod_{h\inI}\!\mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$}$ consisting of {\em coherent sequences}, i.e.\ of those generalized sequences for which \begin{equation} \ff {h'}h \circ \psi(h') = \psi(h) \labl j for all $h,h'\!\in\!\mbox{$I$}$ with $h\preceq h'$, is isomorphic to the projective limit. More precisely, \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}\ is isomorphic to \mbox{$\cal L$}\ as a set, and the morphisms $\ef h$ are the projections to the components, i.e.\ \begin{equation} \ef h(\psi):=\psi(h) \,. \end{equation} For the projective system introduced in subsection \ref{s.ps} where \mbox{${\cal C}$}\ is the (small) category ${\cal F}\!us\mbox{\scriptsize(}\mbox{$\mathfrak g$}\mbox{\scriptsize)}$, the projective limit is clearly {\em not\/} contained in the original category, because no object \mbox{$^{\scriptscriptstyle(i)\!}{\cal R}$}\ of ${\cal F}\!us\mbox{\scriptsize(}\mbox{$\mathfrak g$}\mbox{\scriptsize)}$\ can possess morphisms to {\em all\/} objects \mbox{$^{\scriptscriptstyle(j)\!}{\cal R}$}. In order to identify nevertheless a projective limit associated to the projective system defined by \erf', it is therefore necessary to consider the set \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}\ of coherent sequences. In accordance with the remarks above, for definiteness from now on we will simply refer to \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}\ as `the' projective limit of the system \erf1 of WZW fusion rings. \subsection{Properties of \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}} Let us list a few simple properties of the projective limit \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}. First, \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}\ is a ring over ${\dl Z}$. The product $\psi_1\star \psi_2$ in \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}\ is defined pointwise, i.e.\ by the requirement that \begin{equation} (\psi_1\star \psi_2)(h):= \psi_1(h)\star \psi_2(h) \labl{pcp} for all $h\iN\II$. This definition makes sense, i.e.\ for all $\psi_1,\psi_2\!\in\!\mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}$ also their product is in \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}, because \begin{equation} \begin{array}{ll} \ff{h'}h\circ(\psi_1\star \psi_2)(h') \!\!&= (\ff{h'}h\circ\psi_1(h'))\star (\ff{h'}h\circ\psi_2(h')) \\[1.9mm]& = \psi_1(h) \star \psi_2(h) = (\psi_1\star \psi_2)(h) \,; \end{array}\end{equation} here in the first line the morphism property of the maps $\ff{h'}h$ is used. {}From the definition \erf{pcp} it is clear that the product of \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}\ is commutative and associative, and that \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}\ is unital, with the unit element being the element $\psi_\circ\!\in\!\mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}$ that satisfies \begin{equation} \psi_\circ(h)={}^{\sssh\!}\varphi^{}_{\!\rho} \end{equation} for all $h\iN\II$. Second, a conjugation $\psi\mapsto \psi^+$ can be defined on \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}\ by setting \begin{equation} \psi^+(h):=(\psi(h))^+ \end{equation} for all $h\iN\II$. The conjugation ${}^{\sssh\!}\varphi^{}\mapsto({}^{\sssh\!}\varphi)^+$ on the rings \mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$}\ commutes with the projections $\ff{h'}h$. As a consequence, indeed $\psi^+\!\in\!\mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}$ whenever $\psi\!\in\!\mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}$, conjugation is an involutive au\-to\-morphism of \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}, and the unit element $\psi_\circ$ is self-conjugate. In \secref{binf} we will construct a countable basis \mbox{$^{\scriptscriptstyle(\infty)\!\!}{\cal B}$}\ of the ring \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}; this basis contains in particular the unit element $\psi_\circ$. For any $\psi\!\in\!\mbox{$^{\scriptscriptstyle(\infty)\!\!}{\cal B}$}$ and any $h\iN\II$, $\psi(h)$ is either zero or, up to possibly a sign, an element of the basis \mbox{$^{\scriptscriptstyle(h)}\!{\cal B}$}\ of \mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$}. Also, while by construction the structure constants in the basis \mbox{$^{\scriptscriptstyle(\infty)\!\!}{\cal B}$}\ are integers, there seems to be no reason why they should be non-negative. Accordingly, an interpretation of the limit \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}\ as the representation\ ring of some underlying algebra ic structure is even less obvious than in the case of the fusion rings \mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$}.\,% \futnote{The latter can e.g.\ be regarded as the representation\ rings of the `quantum symmetry' of the associated WZW theories. However, so far there is no agreement on the precise nature of those quantum symmetries.} In particular, in \secref{gb} we will see that \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}\ does not coincide with the representation\ ring \mbox{$\overline{\mbox{${\cal R}$}}$}\ of the simple Lie algebra\ $\mbox{$\bar{\mathfrak g}$}\subset\mbox{$\mathfrak g$}$, but rather that it contains \mbox{$\overline{\mbox{${\cal R}$}}$}\ as a tiny proper subring. As it turns out, the fusion product of two elements of \mbox{$^{\scriptscriptstyle(\infty)\!\!}{\cal B}$}\ is generically {\em not\/} a finite linear combination of elements of \mbox{$^{\scriptscriptstyle(\infty)\!\!}{\cal B}$}, or in other words, \mbox{$^{\scriptscriptstyle(\infty)\!\!}{\cal B}$}\ does not constitute an ordinary basis of \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}. Rather, it must be regarded as a topological basis. For this interpretation to make sense, a suitable topology on \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}\ must be defined. This will be achieved in the next subsection. \subsection{The \topo\ of \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}}\label{topo} The fusion rings \mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$}\ can be considered as topological spaces by simply endowing them with the discrete topology, i.e.\ by declaring every subset to be open (and hence also every subset to be closed). The projective limit \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}\ then becomes a topological space in a natural manner, namely by defining its topology as the coarsest topology in which all projections $\ef h\!:\; \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}\to\mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$}$ are continuous; this will be called the {\em\topo\/} on \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}. \futnot{This is the natural adaptation of the product topology on the direct product $\prod_{h\inI}\mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$}$.} More explicitly, the \topo\ on \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}\ is described by the property that any open set in \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}\ is an (arbitrary, i.e.\ not necessarily finite nor even countable) union of elements of \begin{equation} \mbox{$\Omega$}:= \{ \efm h(M) \mid h\!\in\!\mbox{$I$},\; M\subseteq\mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$} \} \,, \labl{Om} i.e.\ of the set of all pre-images of all sets in any of the fusion rings \mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$}.\\ Note that we need not require to take also finite intersections of these pre-images. This is because \mbox{$\Omega$}\ is closed under taking finite intersections, as can be seen as follows. Let $\omega_i\!\in\!\mbox{$\Omega$}$ for $i=1,2,...\,,N$; by definition, each of the $\omega_i$ can be written as $\omega_i=\efm{h_i}(M_i)$ for some heights $h_i\iN\II$ and some subsets $M_i\subseteq\mbox{$^{\scriptscriptstyle(h_i)\!}{\cal R}$}$. Denote then by $h$ the smallest common multiple of the $h_i$ for $i=1,2,...\,,N$. Because of \erf I we have $\ef{h_i}=\ff h{h_i}\!\circ\!\ef h$, so that \begin{equation} \efm{h_i}(M_i)=\efm h(\ffm h{h_i}(M_i))=\efm h(\tilde M_i) \,, \end{equation} where for all $i=1,2,...\,,N$ the sets $\tilde M_i:=\ffm h{h_i}(M_i)$ are subsets of \mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$}. Because of $\bigcap_{i=1}^N\tilde M_i\subseteq \mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$}$, it thus follows that \begin{equation} \bigcap_{i=1}^N \omega_i =\bigcap_{i=1}^N \efm h(\tilde M_i) = \efm h\mbox{\large[}\bigcap_{i=1}^N \tilde M_i\mbox{\large]} \end{equation} is an element of the set \mbox{$\Omega$}\ \erf{Om}. Thus \mbox{$\Omega$}\ is closed under taking finite intersections, as claimed.\\ As a consequence of this property of \mbox{$\Omega$}, in particular any non-empty open set in \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}\ contains a subset which is of the form $\efm h(M)$ for some $h\iN\II$ and some $M\subseteq\mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$}$; for later reference, we call this fact the `pre-image property' of the non-empty open sets in \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}. Note that the \topo\ on \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}\ is {\em not\/} the discrete one, but finer. To see this, suppose the \topo\ were the discrete one. Then for any $\psi\!\in\! \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}$ the one-element set $\{\psi\}$ would be open and hence a union of sets in $\Omega$ \erf{Om}; but as $\{\psi\}$ just contains one single element, this means that it even has to belong itself to $\Omega$. This in turn means that there would exist $h\iN\II$ and $M\subseteq\mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$}$ such that $\{\psi\}=f_h^{-1}(M)$, and hence simply $M=\{\psi(h)\}$. This, however, would imply that each element $\psi\!\in\!\mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}$ would already be determined uniquely by the value $\psi(h)$ for a single height $h$. {}From the explicit description of \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}\ as a space of coherent generalized sequences, it follows that this is definitely not true. Thus the assumption that the \topo\ is the discrete one leads to a contradiction. Whenever two elements $\psi,\psi'\!\in\!\mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}$ are distinct, there exists some height $h\iN\II$ such that $\psi(h)\ne\psi'(h)$. The open subsets $\omega:=f_h^{-1}(\{\psi(h)\})$ and $\omega':=f_h^{-1}(\{\psi'(h)\})$ then satisfy $\psi\!\in\!\omega$ and $\psi'\!\in\!\omega'$ as well as $\omega\cap\omega'=\emptyset$. This means that when endowed with the \topo, \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}\ is a Hausdorff space. \Sect{A distinguished basis \mbox{$^{\scriptscriptstyle(\infty)\!\!}{\cal B}$}\ of \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}}{binf} In this section we construct a (topological) basis \mbox{$^{\scriptscriptstyle(\infty)\!\!}{\cal B}$}\ of the projective limit \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}\ of WZW fusion rings. \subsection{A linearly independent subset of \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}}\label{s1} We start by defining the subset $\mbox{$^{\scriptscriptstyle(\infty)\!\!}{\cal B}$}\subset\mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}$ as the set of all those elements $\psi\!\in\!\mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}$ which for every $h\iN\II$ satisfy \begin{equation} \psi(h)= \epsilon_h\cdot{}^{\sssh\!}\varphi^{}_a \labl{bd} for some \begin{equation} a\!\in\!\mbox{$_{}^{\scriptscriptstyle(h)\!\!}P$} \qquad{\rm and}\qquad \epsilon_h\!\in\!\{0,\pm1\} \end{equation} (i.e.\ for each height $h$ the fusion ring element $\psi(h)\!\in\!\mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$}$ is either zero or, up to a sign, an element of the distinguished basis \mbox{$^{\scriptscriptstyle(h)}\!{\cal B}$}), and for which in addition not all of the prefactors $\epsilon_h$ vanish and $\epsilon_h=1$ for the smallest $h\iN\II$ for which $\epsilon_h\ne0$. The latter requirement ensures that $-\psi\not\in\mbox{$^{\scriptscriptstyle(\infty)\!\!}{\cal B}$}$ for all $\psi\!\in\!\mbox{$^{\scriptscriptstyle(\infty)\!\!}{\cal B}$}$. Note that at this point we cannot tell yet whether the set \mbox{$^{\scriptscriptstyle(\infty)\!\!}{\cal B}$}\ is large enough to generate the whole ring \mbox{$^{\scriptscriptstyle(\infty)\!\!}{\cal B}$}; in fact, it is even not yet clear whether \mbox{$^{\scriptscriptstyle(\infty)\!\!}{\cal B}$}\ is non-empty. These issues will be dealt with in subsections \ref{s2} to \ref{sinw} below, where we will in particular see that the set \mbox{$^{\scriptscriptstyle(\infty)\!\!}{\cal B}$}\ is countably infinite. However, what we already can see is that the set \mbox{$^{\scriptscriptstyle(\infty)\!\!}{\cal B}$}\ is linearly independent. To prove this, consider any set of finitely many distinct elements $\psi_i$, $i=1,2,...\,,N$ of \mbox{$^{\scriptscriptstyle(\infty)\!\!}{\cal B}$}. We first show that to any pair $i,j\in\{1,2,...\,,N\}$ there exists a height $h_{ij}\iN\II$ such that\\[.2em] \mbox{~~~~~}(i)~~$\psi_i(h_{ij})\ne 0$ \,and\, $\psi_j(h_{ij})\ne 0$ \,\ \ and\\[.2em] \mbox{~~~~~}(ii)~$\psi_i(h_{ij})\ne \pm\psi_j(h_{ij})$\,.\\[.2em] To see this, assume that the statement is wrong, i.e.\ that for each height $h$ either one of the elements $\psi_i(h)$ and $\psi_j(h)$ of \mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$}\ vanishes, or one has $\psi_i(h)= \pm\psi_j(h)$. Now because of $\psi_i\ne0$ and $\psi_j\ne0$ there exists heights $h_i$ and $h_j$ with $\psi_i(h_i)\ne0$ and $\psi_j(h_j)\ne0$. This implies that also $\psi_i(\tilde h_{ij})\ne 0$ and $\psi_j(\tilde h_{ij})\ne 0$ for $\tilde h_{ij}:=h_ih_j$. By our assumption it then follows that $\psi_j(\tilde h_{ij})= \pm \psi_i(\tilde h_{ij})$, which in turn implies that $\psi_j(h_i)= \pm \psi_i(h_i)\ne0$. Now this conclusion actually extends to arbitrary heights $h$. Namely, from the previous result we know that for any $h$ the elements $\psi_i(h\tilde h_{ij})$ and $\psi_j(h\tilde h_{ij})$ must both be non-zero. By our assumption this implies that $\psi_j(h \tilde h_{ij})=\pm\psi_i(h \tilde h_{ij})$. Projecting this equation down to the height $h$, it follows that $\psi_j(h)=\pm\psi_i(h)$. Since $h$ was arbitrary, it follows that in fact $\psi_j= \pm\psi_i$, and hence (as $-\psi_i$ is not in \mbox{$^{\scriptscriptstyle(\infty)\!\!}{\cal B}$}) that $\psi_j=\psi_i$. This is in contradiction to the requirement that all $\psi_i$ should be distinct. Thus our assumption must be wrong, which proves that (i) and (ii) are fulfilled. Applying now the properties (i) and (ii) for any pair $i,j\in\{1,2,...\,,N\}$ with $i\ne j$, it follows that at the height $h:=\prod_{i,j;i<j}h_{ij}$ \futnot{It is already sufficient to take $h$ as the s.c.m.\ of the $h_{ij}$.} we have\\[.2em] \mbox{~~~~~}(i)~~$\psi_i(h)\ne 0$ \,for all\, $i=1,2,...\,,N$ \,\ \ and\\[.2em] \mbox{~~~~~}(ii)~$\psi_i(h)\ne \pm\psi_j(h)$ \,for all $i,j\in\{1,2,...\,,N\}$, $i\ne j$\,.\\[.2em] Thus all the elements $\psi_i(h)$ of \mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$}\ are distinct and, up to sign, elements of the distinguished basis \mbox{$^{\scriptscriptstyle(h)}\!{\cal B}$}. This implies in particular that the only solution of the equation $\sum_{i=1}^N\xi_i\psi_i(h)=0$ is $\xi_i=0$ for $i=1,2,...\,,N$, which in turn shows that also the equation $\sum_{i=1}^N\xi_i\psi_i=0$ has only this solution. Thus, as claimed, the $\psi_i$ are linearly independent elements of \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}. \subsection{\mbox{$^{\scriptscriptstyle(\infty)\!\!}{\cal B}$}\ generates all of \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}}\label{s2} Next we claim that the set \mbox{$^{\scriptscriptstyle(\infty)\!\!}{\cal B}$}\ spans \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}\ in the sense that the closure (in the \topo) of the linear span of \mbox{$^{\scriptscriptstyle(\infty)\!\!}{\cal B}$}\ in \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}, i.e.\ of the set \begin{equation} \mbox{$\langle^{\scriptscriptstyle(\infty)\!\!}{\cal B}\rangle$} \equiv {\rm span}_{\dl Z}^{}(\mbox{$^{\scriptscriptstyle(\infty)\!\!}{\cal B}$}) \end{equation} of finite ${\dl Z}$-linear combinations of elements of \mbox{$^{\scriptscriptstyle(\infty)\!\!}{\cal B}$}, is already all of \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}. To prove this, assume that the statement is wrong, or in other words, that the set \begin{equation} {\cal S} := \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}\,\setminus\,\mbox{$\overline{\langle^{\scriptscriptstyle(\infty)\!\!}{\cal B \labl{mw} is non-empty. By definition, the set ${\cal S}$ is open, and hence because of the pre-image property\ it contains a subset ${\cal M}\subseteq{\cal S}$ of the form ${\cal M}=\efm h(M)$ for some $h\iN\II$ and some $M\subseteq\mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$}$. Further, as an immediate consequence of the construction that we will present in the subsections \ref{siw} and \ref{sinw}, for each $a\!\in\!\mbox{$_{}^{\scriptscriptstyle(h)\!\!}P$}$ there exists an element (in fact, infinitely many elements) $\psi_a\!\in\!\mbox{$^{\scriptscriptstyle(\infty)\!\!}{\cal B}$}$ such that $\ef h(\psi_a)={}^{\sssh\!}\varphi^{}_a$ (namely, we need to prescribe the value of $\psi_a(p)$ only for the finitely many prime factors $p$ of $h$). Now choose some $y\!\in\! M$, decompose it with respect to the basis \mbox{$^{\scriptscriptstyle(h)}\!{\cal B}$}\ of \mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$}, i.e. $y=\sumpH a n_a\,{}^{\sssh\!}\varphi^{}_a$ with $n_a\!\in\!{\dl Z}$, and define $\eta:=\sumpH a n_a\,\psi_a$. Then, on one hand, by construction we have $\ef h(\eta)=y$, i.e.\ $\eta\!\in\!{\cal M}$, and hence $\eta\in{\cal S}$, while on the other hand $\eta$ is a finite linear combination of elements of \mbox{$^{\scriptscriptstyle(\infty)\!\!}{\cal B}$}\ (since $y$ is a finite linear combination of elements of \mbox{$^{\scriptscriptstyle(h)}\!{\cal B}$}), and hence $\eta\in\mbox{$\langle^{\scriptscriptstyle(\infty)\!\!}{\cal B}\rangle$}\subseteq\mbox{$\overline{\langle^{\scriptscriptstyle(\infty)\!\!}{\cal B$. By the definition \erf{mw} of ${\cal S}$, this is a contradiction, and hence our assumption must be wrong. Together with the result of the previous subsection we thus see that \mbox{$^{\scriptscriptstyle(\infty)\!\!}{\cal B}$}\ is a (topological) basis of \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}. \subsection{Distinguished sequences of integral weights}\label{siw} We will now construct all elements of the projective limit \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}\ which belong to the subset \mbox{$^{\scriptscriptstyle(\infty)\!\!}{\cal B}$}\ as introduced in subsection \ref{s1}. These are obtained as generalized sequences $\psi$ satisfying both \erf j and the defining relation \erf{bd} of \mbox{$^{\scriptscriptstyle(\infty)\!\!}{\cal B}$}. More specifically, we construct sequences ${({a_h^{}})}_{h\in\II}$ of labels ${a_h^{}}\!\in\!\mbox{$_{}^{\scriptscriptstyle(h)\!\!}P$}$ and associated signs $\eta({a_h^{}})$ such that all those $\psi$ which are of the form \begin{equation} \psi(h)=\eta({a_h^{}})\,{}^{\sssh\!}\varphi^{}_{{a_h^{}}} \labl Q belong to the subset $\mbox{$^{\scriptscriptstyle(\infty)\!\!}{\cal B}$}\subset\mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}$. When applied to \erf Q, the requirement \erf j amounts to \begin{equation} \ff{\el h}h({}^{\ssslh\!}\varphi^{}_{a_{\ell h}^{}}) = \eta({a_{\ell h}^{}})\eta({a_h^{}})\cdot{}^{\sssh\!}\varphi^{}_{a_h^{}} \,, \labl k which in view of the definition \erf' of \ff{\el h}h\ is equivalent to \begin{equation} {a_{\ell h}^{}}=w({a_h^{}}) \qquad{\rm for\ some}\ w\!\in\!\mbox{$^{\sssh\!}W$} \labl m and \begin{equation} \eta({a_{\ell h}^{}})\eta({a_h^{}})=\epsilon_{\el}({a_{\ell h}^{}}) \,. \labl M To start the construction of the elements of \mbox{$^{\scriptscriptstyle(\infty)\!\!}{\cal B}$}, we first concentrate our attention to integral weights of the height $h$ theory which are not necessarily integrable and which are considered as defined only modulo $h\mbox{$\overline L^{\scriptscriptstyle\vee}$}$; we denote these weights by ${b_h^{}}$. Suppose then that we prescribe for each prime $p$ such a weight ${b_p^{}}$ and that these weights satisfy in addition the restriction that for any two primes $p,\,p'$ they differ by an element of the coroot lattice, \begin{equation} {b_p^{}}-{b_{p'}^{}} \in \mbox{$\overline L^{\scriptscriptstyle\vee}$} \,. \labl A We claim that there then exists a sequence ${({b_h^{}})}_{h\in\II}$ which for prime heights takes the prescribed values ${b_p^{}}$ and for which the relation \begin{equation} {b_{h'}^{}} = {b_h^{}} \ {\rm mod}\; h\mbox{$\overline L^{\scriptscriptstyle\vee}$} \labl F holds for all $h,\,h'$ with $h\preceq h'$. To prove this assertion, we display such a sequence explicitly. To this end, let \begin{equation} h=:\prod_{\scriptstyle j\atop \scriptstyle p_j|h}{p_j^{n_j}} \labl G denote the decomposition of $h$ into prime factors, and define \begin{equation} {h_i}:=\frac{h}{p_i^{n_i}} \labl H and \begin{equation} \invmod\hi{p_i^{n_i}}:= ({h_i})^{-1}_{} \ {\rm mod}\; {p_i^{n_i}} \,. \labl J Then we set \begin{equation} {b_h^{}} := {b_{p_1^{}}^{}} + \sumieh {h_i}\,\invmod\hi{p_i^{n_i}}\,({b_{p_i^{}}^{}}-{b_{p_1^{}}^{}}) \,. \labl{bh} Recall that ${b_h^{}}$ is defined only modulo $h\mbox{$\overline L^{\scriptscriptstyle\vee}$}$. In \erf{bh} $p_1$ is any of the prime divisors of $h$; it has been singled out only in order to make the formula for ${b_h^{}}$ to look as simple as possible, and in fact ${b_h^{}}$ does not depend (modulo $h\mbox{$\overline L^{\scriptscriptstyle\vee}$}$) on the choice of $p_1$. To see this, let ${b_h^{\scriptscriptstyle(2)}}$ denote the number obtained analogously as in \erf{bh}, but with $p_1$ replaced by some other prime factor $p_2$ of $h$. Then \begin{equation} \begin{array}{l} {b_h^{}}-{b_h^{\scriptscriptstyle(2)}}={b_{p_1^{}}^{}}-{b_{p_2^{}}^{}} + \sumiezh {h_i}\,\invmod\hi{p_i^{n_i}}\,({b_{p_2^{}}^{}}-{b_{p_1^{}}^{}}) \\[1.5mm] \hsp{19.1} + {h_2}\,\invmod\hz{p_2^{n_2}}\,({b_{p_2^{}}^{}}-{b_{p_1^{}}^{}}) - {h_1}\,\invmod\he{p_1^{n_1}}\,({b_{p_1^{}}^{}}-{b_{p_2^{}}^{}}) \,. \end{array}\end{equation} Using the fact that ${h_i}$ is divisible by ${p_j^{n_j}}$ for all primes $p_j$ dividing $h$ except for $j=i$, and that ${h_i}\invmod\hi{p_i^{n_i}}=1\ {\rm mod}\;{p_i^{n_i}}$, it is easily checked that the right hand side of this expression vanishes modulo ${p_j^{n_j}}\mbox{$\overline L^{\scriptscriptstyle\vee}$}$ for all $p_j$ dividing $h$, and hence, using \erf A, also vanishes modulo $h\mbox{$\overline L^{\scriptscriptstyle\vee}$}$. To establish the coherence property \erf F, we now consider two heights $h,\,h'$ such that $h|h'$. Then we set \begin{equation} h'=:\prod_{\scriptstyle j\atop \scriptstyle p_j|h'}{p_j^{n'_j}} \end{equation} and ${h_{i}'}:=h'/{p_i^{n'_i}}$, and without loss of generality we can assume that $p_1$ divides $h$ as well as $h'$. By the definition \erf{bh} we then have \begin{equation} {b_{h'}^{}}-{b_h^{}}= \sumieh \mbox{\Large\{} {h_{i}'}\,\invmod\hip{p_i^{n'_i}} - {h_i}\,\invmod\hi{p_i^{n_i}} \mbox{\Large\}} \,({b_{p_i^{}}^{}}-{b_{p_1^{}}^{}}) +\!\!\! \sumihh {h_{i}'}\,\invmod\hip{p_i^{n'_i}} \,({b_{p_i^{}}^{}}-{b_{p_1^{}}^{}}) \,; \end{equation} again it is straightforward to verify that this vanishes modulo ${p_j^{n_j}}\mbox{$\overline L^{\scriptscriptstyle\vee}$}$ for all $p_j$ dividing $h$. This shows that the property \erf F is satisfied for the sequence defined by \erf{bh} as claimed. Next we note that we did not require that the prescribed values ${b_p^{}}$ lie on the Weyl orbit of an integrable weight at height $p$, but rather they may also lie on the boundary of some Weyl chamber of \mbox{$^{\sss(p)\!}W$}. However, if ${b_p^{}}$ does belong to the Weyl orbit of an integrable weight, then also each weight ${b_h^{}}$ with $p|h$ is on the Weyl orbit of an integrable weight at height $h$. Namely, because of the property \erf F we have in particular \begin{equation} {b_h^{}} = {b_p^{}} + p^n\,\beta \end{equation} for some $\beta\!\in\!\mbox{$\overline L^{\scriptscriptstyle\vee}$}$. Hence, assuming that ${b_h^{}}$ is left invariant by some $w\!\in\!\mbox{$^{\sssh\!}W$}$, i.e.\ that ${b_h^{}}=w({b_h^{}})\equiv\overline w({b_h^{}})+h\gamma$ for some element $\gamma$ of the coroot lattice, it follows that \begin{equation} \begin{array}{l} \overline w({b_p^{}})= \overline w({b_h^{}}-p^n\beta) = \overline w({b_h^{}})-p^n\,\overline w(\beta) \\[1.5mm] \hsp{9.9} = {b_h^{}} - h\gamma -p^n\,\overline w(\beta) = {b_p^{}} +p^n\,(\beta-\overline w(\beta))- h\gamma \,. \end{array}\end{equation} Since by assumption the only element of \mbox{$^{\sss(p)\!}W$}\ which leaves the weight ${b_p^{}}$ invariant is the identity, it follows that $\overline w={\sf id}$ and $\gamma=0$, implying that also the only element of \mbox{$^{\sssh\!}W$}\ that leaves ${b_h^{}}$ invariant is the identity, which is equivalent to the claimed property. Our next task is to investigate to what extent the sequence ${({b_h^{}})}_{h\in\II}$ is characterized by the prescribed values ${b_p^{}}$ at prime heights and by the requirement \erf F. To this end let ${({\tilde b_h^{}})}_{h\in\II}$ be another such sequence, i.e.\ a sequence such that ${\tilde b_p^{}}={b_p^{}}$ for all primes $p$ and ${\tilde b_{h'}^{}}-{\tilde b_h^{}}\in h\mbox{$\overline L^{\scriptscriptstyle\vee}$}$ for $h|h'$. First we observe that for $h$ and $h'$ coprime, the properties \begin{equation} {\tilde b_{hh'}^{}} = {\tilde b_h^{}} \ {\rm mod}\; h\mbox{$\overline L^{\scriptscriptstyle\vee}$} \qquad{\rm and}\qquad {\tilde b_{hh'}^{}} = {\tilde b_{h'}^{}} \ {\rm mod}\; h'\mbox{$\overline L^{\scriptscriptstyle\vee}$} \end{equation} fix ${\tilde b_{hh'}^{}}$ already uniquely (modulo $hh'\mbox{$\overline L^{\scriptscriptstyle\vee}$}$), so that the whole freedom is parametrized by the freedom in the choice of ${\tilde b_h^{}}$ at heights which are a prime power. Concerning the latter freedom, we claim that for any prime $p$ there is a sequence of elements $\beta_p^{{\scriptscriptstyle(}j{\scriptscriptstyle)}}$ of the coroot lattice $\mbox{$\overline L^{\scriptscriptstyle\vee}$}$ which are defined modulo $p\mbox{$\overline L^{\scriptscriptstyle\vee}$}$ such that the most general choice of ${\tilde b_{p^n}^{}}$ reads \begin{equation} {\tilde b_{p^n}^{}} = {b_{p^n}^{}} + \sumne j \beta_p^{{\scriptscriptstyle(}j{\scriptscriptstyle)}}\,p^j \labl{bb} with ${b_{p^n}^{}}$ defined according to \erf{bh}, i.e.\ simply ${b_{p^n}^{}}={b_p^{}}$. This statement is proven by induction. For $n=1$ it is trivially fulfilled. Further, assuming that \erf{bb} is satisfied for some $n\ge1$ and setting $\gamma:={\tilde b_{p^{n+1}}^{}}-{b_{p^{n+1}}^{}}$ (defined modulo $p^{n+1}\mbox{$\overline L^{\scriptscriptstyle\vee}$}$), one has \begin{equation} {\tilde b_{p^{n+1}}^{}}-{\tilde b_{p^n}^{}}= \gamma-\sumne j \beta_p^{{\scriptscriptstyle(}j{\scriptscriptstyle)}}\,p^j \,. \end{equation} By the required properties of the sequence $({\tilde b_h^{}})$, the left hand side of this formula must vanish modulo $p^n\mbox{$\overline L^{\scriptscriptstyle\vee}$}$, and hence we have \begin{equation} \gamma = \beta_p^{{\scriptscriptstyle(}n{\scriptscriptstyle)}}\,p^n + \sumne j \beta_p^{{\scriptscriptstyle(}j{\scriptscriptstyle)}}\,p^j = \sumn j \beta_p^{{\scriptscriptstyle(}j{\scriptscriptstyle)}}\,p^j \end{equation} for some $\beta_p^{{\scriptscriptstyle(}n{\scriptscriptstyle)}}\!\in\!\mbox{$\overline L^{\scriptscriptstyle\vee}$}$. This shows that ${\tilde b_{p^{n+1}}^{}}$ is again of the form described by \erf{bb}; furthermore, as $\gamma$ is defined modulo $p^{n+1}\mbox{$\overline L^{\scriptscriptstyle\vee}$}$, $\beta_p^{{\scriptscriptstyle(}n{\scriptscriptstyle)}}$ is defined modulo $p\mbox{$\overline L^{\scriptscriptstyle\vee}$}$ as required, and hence the proof of the formula \erf{bb} is completed. With these results we are now in a position to give a rather explicit description of the allowed sequences ${\tilde b_h^{}}$. Namely, we can parametrize the general form of ${\tilde b_h^{}}$ in terms of the freedom in ${\tilde b_{p^n}^{}}$ according to \begin{equation} \begin{array}{l} \hsp{-3}{\tilde b_h^{}}= {\tilde b_{p_1^{n_1}}^{}} + \sumieh{h_i}\,\invmod\hi{p_i^{n_i}}\,({\tilde b_{p_i^{n_i}}^{}}-{\tilde b_{p_1^{n_1}}^{}}) \\[1.5mm] \hsp{.7} = {b_{p_1^{}}^{}} + \sumnee j \beta_{p_1}^{{\scriptscriptstyle(}j{\scriptscriptstyle)}}\,p_1^j + \sumieh {h_i}\,\invmod\hi{p_i^{n_i}}\,({b_{p_i^{}}^{}}-{b_{p_1^{}}^{}}) + \sumieh {h_i}\,\invmod\hi{p_i^{n_i}}\,\mbox{\large[} \sumnie j \beta_{p_i}^{{\scriptscriptstyle(}j{\scriptscriptstyle)}}\,p_i^j - \sumnee j \beta_{p_1}^{{\scriptscriptstyle(}j{\scriptscriptstyle)}}\,p_1^j \mbox{\large]} \\[1.5mm] \hsp{.7} = {b_h^{}} + \sumieh {h_i}\,\invmod\hi{p_i^{n_i}}\,\mbox{\large[} \sumnie j \beta_{p_i}^{{\scriptscriptstyle(}j{\scriptscriptstyle)}}\,p_i^j - \sumnee j \beta_{p_1}^{{\scriptscriptstyle(}j{\scriptscriptstyle)}}\,p_1^j \mbox{\large]} + \sumnee j \beta_{p_1}^{{\scriptscriptstyle(}j{\scriptscriptstyle)}}\,p_1^j \,. \end{array} \labl X Further, any such sequence fulfills the consistency requirement that ${\tilde b_{h'}^{}}-{\tilde b_h^{}}\in h\mbox{$\overline L^{\scriptscriptstyle\vee}$}$ for heights $h,\,h'$ with $h|h'$. Namely, in this case the formula \erf X yields \begin{equation} \begin{array}{l} \hsp{-3.8} {\tilde b_{h'}^{}}-{\tilde b_h^{}} = \sumieh \mbox{\Large\{} {h_{i}'}\,\invmod\hip{p_i^{n'_i}}\, \mbox{\large[} \sumnipe j \beta_{p_i}^{{\scriptscriptstyle(}j{\scriptscriptstyle)}}\,p_i^j - \sumnepe j \beta_{p_1}^{{\scriptscriptstyle(}j{\scriptscriptstyle)}}\,p_1^j \mbox{\large]} - {h_i}\,\invmod\hi{p_i^{n_i}}\, \mbox{\large[} \sumnie j \beta_{p_i}^{{\scriptscriptstyle(}j{\scriptscriptstyle)}}\,p_i^j - \sumnee j \beta_{p_1}^{{\scriptscriptstyle(}j{\scriptscriptstyle)}}\,p_1^j \mbox{\large]} \mbox{\Large\}} \\[1.5mm] \hsp{16.4} + \sumihh {h_{i}'}\,\invmod\hip{p_i^{n'_i}}\, \mbox{\large[} \sumnipe j \beta_{p_i}^{{\scriptscriptstyle(}j{\scriptscriptstyle)}}\,p_i^j - \sumnepe j \beta_{p_1}^{{\scriptscriptstyle(}j{\scriptscriptstyle)}}\,p_1^j \mbox{\large]} + \sumneee j \beta_{p_1}^{{\scriptscriptstyle(}j{\scriptscriptstyle)}}\,p_1^j \ \ {\rm mod}\; h\mbox{$\overline L^{\scriptscriptstyle\vee}$} \,. \end{array} \labl Y Once more one can easily check that this expression vanishes modulo ${p_j^{n_j}}\mbox{$\overline L^{\scriptscriptstyle\vee}$}$ for all primes $p_j$ that divide $h$, and hence vanishes modulo $h\mbox{$\overline L^{\scriptscriptstyle\vee}$}$. Thus the consistency requirement indeed is satisfied. \subsection{Distinguished sequences of integrable weights and the basis \mbox{$^{\scriptscriptstyle(\infty)\!\!}{\cal B}$}} \label{sinw} What we have achieved so far is a characterization of all sequences ${({b_h^{}})}_{h\in\II}$ of integral weights defined modulo $h\mbox{$\overline L^{\scriptscriptstyle\vee}$}$ that satisfy \erf F. We now use this result to construct sequences ${({a_h^{}})}_{h\in\II}$ of highest weights which satisfy the requirement \erf m and of which infinitely many are {\em integrable\/} weights, with all non-integrable weights being equal to zero. We start by prescribing integrable weights ${a_p^{}}\!\in\!\mbox{$_{}^{\scriptscriptstyle(p)\!\!}P$}$ for all primes $p$ with $p\ge\mbox{$g_{}^{\scriptscriptstyle\vee}$}$, and set ${b_p^{}}={a_p^{}}$ for $p\ge\mbox{$g_{}^{\scriptscriptstyle\vee}$}$, while for all primes $p<\mbox{$g_{}^{\scriptscriptstyle\vee}$}$ we choose arbitrary weights ${b_p^{}}$ (which are necessarily non-integrable). Next we employ the previous results to find the sequences ${({b_h^{}})}_{h\in\II}$. Finally we define ${a_h^{}}$ for any arbitrary height $h$ as follows. If ${b_h^{}}$ lies on the boundary of a Weyl chamber with respect to \mbox{$^{\sssh\!}W$}, then we set ${a_h^{}}=0$. Otherwise there are a unique element $\overline w_h\in\overline W$ and a unique \futnote{To be precise, because the weights ${a_h^{}}$ are defined only modulo $h\mbox{$\overline L^{\scriptscriptstyle\vee}$} $, $\gamma_h$ is only unique once a definite representative of the equivalence class of weights that is described by ${a_h^{}}$ is chosen.} element $\gamma_h\!\in\!\mbox{$\overline L^{\scriptscriptstyle\vee}$}$ such that $\overline w_h({b_h^{}})+h\gamma_h$ is integrable, and in this case we set \begin{equation} {a_h^{}}:= \overline w_h({b_h^{}})+h\gamma_h \,. \end{equation} By construction, the weights ${a_h^{}}$ have the following properties. If ${a_{h'}^{}}=0$ for some height $h'$, then ${b_{h'}^{}}$ is on the boundary of a Weyl chamber with respect to \mbox{$^{\sss(h')\!}W$}; for any $h$ dividing $h'$, it then follows from ${b_{h'}^{}}-{b_h^{}}\!\in\! h\mbox{$\overline L^{\scriptscriptstyle\vee}$}$ that ${b_h^{}}$ is on the boundary of a Weyl chamber with respect to \mbox{$^{\sssh\!}W$}, and hence we also have ${a_h^{}}=0$. On the other hand, if ${b_{h'}^{}}$ is equivalent with respect to \mbox{$^{\sss(h')\!}W$}\ to an integrable weight ${a_{h'}^{}}\!\in\!\mbox{$_{}^{\scriptscriptstyle(h')\!\!}P$}$, then it is a fortiori equivalent to ${a_{h'}^{}}$ with respect to the larger group \mbox{$^{\sssh\!}W$}, and then the property ${b_{h'}^{}}-{b_h^{}}\!\in\! h\mbox{$\overline L^{\scriptscriptstyle\vee}$}$ implies that also ${b_h^{}}$ is equivalent with respect to \mbox{$^{\sssh\!}W$}\ to ${a_{h'}^{}}$, and hence that the associated weight ${a_h^{}}$ is integrable at height $h$ and is equivalent with respect to \mbox{$^{\sssh\!}W$}\ to ${a_{h'}^{}}$, too. Thus ${a_h^{}}$ and ${a_{h'}^{}}$ are on the same orbit with respect to \mbox{$^{\sssh\!}W$}\ whenever $h$ divides $h'$, and hence \erf m holds as promised. Note that by construction for all \mbox{$\bar{\mathfrak g}$}\ except $\mbox{$\bar{\mathfrak g}$}=A_1$ the sequences so obtained contain some zero weights. However, any sequence which contains at least one non-zero weight contains in fact infinitely many non-zero (and hence integrable) weights. The final step is now to define \begin{equation} \psi(h):=\eta({a_h^{}})\,{}^{\sssh\!}\varphi^{}_{{a_h^{}}} \labl x as in \erf Q, where ${a_h^{}}$ is as constructed above, and where \begin{equation} \eta({a_h^{}}):= \left\{ \begin{array}{ll} \mbox{sign}(\overline w_h) & {\rm for}\ {a_h^{}}\!\in\!\mbox{$^{\sssh\!}W$}\,, \\[1mm] 0 & {\rm for}\ {a_h^{}}=0 \,. \end{array} \right. \labl{ets} To show that $\psi$ is an element of the projective limit, it only remains to check the property \erf M of the prefactor $\eta({a_h^{}})$. For ${a_h^{}}=0$ \erf M just reads $0=0$ and is trivially satisfied. For ${a_h^{}}\!\in\!\mbox{$_{}^{\scriptscriptstyle(h)\!\!}P$}$, the previous results show that ${a_{\ell h}^{}}=w_{\ell h}^{}\circ w_0\circ w_h^{-1}({a_h^{}})$, where $w_0$ is the Weyl translation relating ${b_h^{}}$ and ${b_{\ell h}^{}}$, so that \begin{equation} \epsilon_{\el}({a_{\ell h}^{}}) = \mbox{sign}(w_{\ell h}^{}\circ w_0\circ w_h^{}) = \mbox{sign}(\overline w_{\ell h}^{})\cdot\mbox{sign}(\overline w_h^{}) \,. \end{equation} In view of the definition \erf{ets} of $\eta({a_h^{}})$, this is precisely the required relation \erf M. We conclude that the basis \mbox{$^{\scriptscriptstyle(\infty)\!\!}{\cal B}$}\ of \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}\ precisely consists of the elements \erf x. In particular, \mbox{$^{\scriptscriptstyle(\infty)\!\!}{\cal B}$}\ is countably infinite. \Sect{The fusion ring of \mbox{$\bar{\mathfrak g}$}}{gb} As already pointed out in the introduction, it is expected that in the limit of infinite level of WZW theories\ somehow the simple Lie algebra\ \mbox{$\bar{\mathfrak g}$}\ which is the horizontal subalgebra\ of \mbox{$\mathfrak g$}\ and its representation\ theory should play a r\^ole. More specifically, one might think that the representation ring \mbox{$\overline{\mbox{${\cal R}$}}$}\ of \mbox{$\bar{\mathfrak g}$}\ emerges. As we will demonstrate below, indeed this ring shows up, but it is only a proper subring of the projective limit \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}\ we constructed, and almost all elements of \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}\ are {\em not\/} contained in the ring \mbox{$\overline{\mbox{${\cal R}$}}$}. Let us describe \mbox{$\overline{\mbox{${\cal R}$}}$}\ and its connection with the category ${\cal F}\!us\mbox{\scriptsize(}\mbox{$\mathfrak g$}\mbox{\scriptsize)}$\ in some detail. \mbox{$\overline{\mbox{${\cal R}$}}$}\ is defined as the ring over ${\dl Z}$\ of all isomorphism classes of finite-dimensional\ \mbox{$\bar{\mathfrak g}$}-representation s, with the ring product the ordinary tensor product of \mbox{$\bar{\mathfrak g}$}-representation s (or, equivalently, the pointwise product of the characters of these representation s). This ring \mbox{$\overline{\mbox{${\cal R}$}}$}\ is a fusion ring with an infinite basis. The elements $\bar\varphi^{}_a$ of a distinguished basis of \mbox{$\overline{\mbox{${\cal R}$}}$}\ are labelled by the (shifted) highest weights of irreducible finite-dimensional\ \mbox{$\bar{\mathfrak g}$}-representation s, i.e.\ by elements of the set \begin{equation} \mbox{$\bar P$} := \{ a\in\mbox{$\overline L^{\rm w}$} \mid 0<a^i\ {\rm for}\, \ i=1,2,...\,,r \} \,. \labl{pb} Now for any $h\!\in\!\mbox{$I$}$ let us define the map $\fb h\!:\ \mbox{$\overline{\mbox{${\cal R}$}}$}\to\mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$}\,$ as follows. If $a\!\in\!\mbox{$\bar P$}$ lies on the boundary of some Weyl chamber with respect to \mbox{$^{\sssh\!}W$}, we set $\fb h(\bar\varphi^{}_a):=0$; otherwise there exist a unique $a'\!\in\!\mbox{$_{}^{\scriptscriptstyle(h)\!\!}P$}$ and a unique $w\!\in\!\mbox{$^{\sssh\!}W$}$ such that $w(a)=a'$, and in this case we set \begin{equation} \fb h(\bar\varphi^{}_a):= \epsilon(a)\cdot {}^{\sssh\!}\varphi^{}_{a'} \labl` with $\epsilon(a)=\mbox{sign}(w)$. As in the case of the maps $\ff{\el h}h$ \erf', we will consider \erf` as covering all cases, i.e.\ set $\epsilon(a)=0$ if $a$ lies on the boundary of a Weyl chamber at height $h$. To analyze the relation between the ring \mbox{$\overline{\mbox{${\cal R}$}}$}\ and the category ${\cal F}\!us\mbox{\scriptsize(}\mbox{$\mathfrak g$}\mbox{\scriptsize)}$, we first recall the expressions \begin{equation} \nb abc = \sum_{\overline w\in\overline W} \mbox{sign}(\overline w)\,\mult b(\overline w(c)-a) \labl{nb} for the Litt\-le\-wood\hy Ri\-chard\-son coefficient s (or tensor product coefficients) of \mbox{$\bar{\mathfrak g}$}\ \cite{raca2,spei} and \begin{equation} \nh abc = \sum_{w\in{}^{\sssh}W} \mbox{sign}(w)\,\mult b(w(c)-a) \labl{nh} for the fusion rule coefficient s, i.e.\ the structure constants of the WZW fusion ring \mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$}\ \cite{KAc3,walt3,fugp2,fuva2}. Here $\mult a(b)$ denotes the multiplicity of the (shifted) weight $b$ in the \mbox{$\bar{\mathfrak g}$}-representation\ with (shifted) highest weight $a$. It will be convenient to extend the validity of \erf{nb} by adopting it as a definition of \nb abc\ for arbitrary (i.e., not necessarily lying in \mbox{$\bar P$}) integral weights $a$ and $c$, and also extend it to arbitrary integral weights $b$ that do not lie on the boundary of any Weyl chamber with respect to $\overline W$ by setting \begin{equation} \mult b(c):=\mbox{sign}(\overline w_b^{})\,\mult{\overline w_b^{}(b)}(c) \,, \end{equation} with $\overline w_b$ the unique element of $\overline W$ such that $\overline w_b(b)\!\in\!\mbox{$\bar P$}$. The multiplicities $\mult a(b)$ are invariant under the Weyl group $\overline W$, i.e.\ $\mult a(\overline w(b))=\mult a(b)$ for all $\overline w\!\in\!\overline W$. As a consequence, the numbers \nb abc\ and \nh abc\ are related by \futnote{In the formulation of \cite{KAc3,walt3,fugp2,fuva2} the factor of $|\overline W|^{-1}$ is absent because there \nb abc\ is taken to be non-zero only if $a,b\!\in\!\mbox{$\bar P$}$.} \begin{equation} \nh abc = \frac1{|\overline W|}\, \sum_{w\in{}^{\sssh}W} \mbox{sign}(w)\,\nb ab{w(c)} \,. \labl B The invariance of $\mult a(b)$ under $\overline W$ also implies that for arbitrary integral weights $a,\,b$ and $c$ the symmetry property \begin{equation} \nb abc = \nb bac \labl{sy} follows from the analogous property of the Litt\-le\-wood\hy Ri\-chard\-son coefficient s with $a,b,c\in\mbox{$\bar P$}$, and that \begin{equation} \begin{array}{l} \nb{\overline w_1(a)}b{\,\ \overline w_2(c)} = \displaystyle\sum_{\overline w\in\overline W} \mbox{sign}(\overline w)\,\mult b(\overline w\,\overline w_2(c)-\overline w_1(a)) \\{}\\[-3mm] \hsp{16.2} = \displaystyle\sum_{\overline w\in\overline W} \mbox{sign}(\overline w)\,\mult b(\overline w_1^{-1}\overline w\:\overline w_2(c)-a) = \mbox{sign}(\overline w_1\overline w_2)\cdot\nb abc \,. \end{array} \end{equation} When combined with the symmetry property \erf{sy}, the latter formula yields \begin{equation} \nb{\overline w_1(a)}{\overline w_2(b)}{\ \ \ \ \ \ \overline w_3(c)} = \mbox{sign}(\overline w_1\overline w_2\overline w_3)\cdot\nb abc \,. \end{equation} To obtain information about the effect of affine Weyl transformations on the labels of \nb abc, we consider an alternating sum over the Weyl group \mbox{$^{\sssh\!}W$}. We have \begin{equation} \begin{array}{l} \displaystyle\sum_{w_2\in{}^{\sssh}W}\!\!\mbox{sign}(w_2)\,\nb{w_1(a)}b{\ \,w_2(c)} = \!\!\displaystyle\sum_{\scriptstyle\overline w,\overline w_2\in\overline W \atop\scriptstyle\beta_2\in\bar L^{\scriptscriptstyle\vee}} \!\!\mbox{sign}(\overline w)\mbox{sign}(\overline w_2) \,\mult b(\overline w\,\overline w_2(c)+h\overline w(\beta_2)-\overline w_1(a)-h\beta_1) \\{}\\[-3mm] \hsp{43.7} = \displaystyle\sum_{\overline w,\overline w_2\in\overline W} \displaystyle\sum_{\beta\in\bar L^{\scriptscriptstyle\vee}} \mbox{sign}(\overline w\,\overline w_2) \,\mult b(\overline w_1^{-1}\overline w\,\overline w_2(c)+h\overline w_1^{-1}\overline w(\beta)-a) \\{}\\[-3mm] \hsp{43.7} = \mbox{sign}(w_1)\cdot\displaystyle\sum_{w_2\in{}^{\sssh}W}\mbox{sign}(w_2)\,\nb ab{w_2(c)} \,. \end{array}\end{equation} Here $\beta:=\beta_2-\overline w^{-1}(\beta_1)$. Together with the symmetry property \erf{sy} it then follows that \begin{equation} \sum_{w_3\in{}^{\sssh}W}\mbox{sign}(w_3)\,\nb{w_1(a)}{w_2(b)}{\ \ \ \ \ w_3(c)} = \mbox{sign}(w_1)\,\mbox{sign}(w_2)\cdot\sum_{w_3\in{}^{\sssh}W}\mbox{sign}(w_3)\,\nb ab{w_3(c)} \labl C for all $w_1,w_2\!\in\!\mbox{$^{\sssh\!}W$}$. We can rewrite this as \begin{equation} \sum_{w\in{}^{\sssh}W}\epsilon_{\el}(A)\epsilon_{\el}(B)\,\mbox{sign}(w)\,\nb{w_A(A)}{w_B(B)} {\ \ \ \ \ \ \ \ \ \ w(c)} = \displaystyle\sum_{w\in{}^{\sssh}W} \mbox{sign}(w)\,\nb AB{w(c)} \, \labl{515} which by interpreting $A$ and $B$ as elements of \mbox{$\bar P$}\ rather than \mbox{$_{}^{\scriptscriptstyle(\el h)\!\!}P$}\ yields, after summation over $c\!\in\!\mbox{$_{}^{\scriptscriptstyle(h)\!\!}P$}$, \begin{equation} \fb h(\bar\varphi^{}_A)\star\fb h(\bar\varphi^{}_B)=\fb h(\bar\varphi^{}_A\star \bar\varphi^{}_B) \,, \end{equation} and hence shows that the maps \fb h defined by \erf` are ring homomorphisms. Now for all $A\!\in\!\mbox{$_{}^{\scriptscriptstyle(\el h)\!\!}P$}$ we have $\ff{\el h}h(\phi_A)=\epsilon_{\el}(A)(\phi_{w_A(A)}) =:\epsilon_{\el}(A)\cdot \varphi^{}_a$. Then owing to \erf B we obtain, after dividing \erf{515} by $|\overline W|$, the relation \begin{equation} \nh{\!\FF{\el h}h(A)}{\FF{\el h}h(B)}{\ \ \ \ \ \ \ \ \ \ \ \ \ c} = \epsilon_{\el}(A)\,\epsilon_{\el}(B)\,\nh abc = \displaystyle\sum_{C:\ \phi_C\in\FF{\el h}h^{-1}(\varphi_c)} \epsilon_{\el}(C)\;\nlh ABC \labl: (on the left hand side, we use the short hand notation $\ff{\el h}h(A)\!\in\!\mbox{$_{}^{\scriptscriptstyle(h)\!\!}P$}$ to indicate the label that corresponds to the element $\epsilon_{\el}(A)\ff{\el h}h(\phi_A)$ of \mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$}). Summation over $c\!\in\!\mbox{$_{}^{\scriptscriptstyle(h)\!\!}P$}$ then yields $\ff{\el h}h(\phi_A)\star\ff{\el h}h(\phi_B)=\ff{\el h}h(\phi_A\star \phi_B)$, so that \erf: is just the homomorphism property of the maps \ff ij which were defined by \erf' in terms of the fusion rule coefficient s. (Thereby we have also obtained an alternative proof of the homomorphism property of those maps.) To investigate further the relation between $\mbox{$\overline{\mbox{${\cal R}$}}$}$ and the projective limit \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}, we introduce the linear mappings \begin{equation} \begin{array}{llll} \jb h:& \mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$}\to\mbox{$\overline{\mbox{${\cal R}$}}$}\,,& {}^{\sssh\!}\varphi^{}_a\mapsto\bar\varphi^{}_a \,, \\[1.6mm] \jj h{h'}:& \mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$}\to\mbox{$^{\scriptscriptstyle(h')\!}{\cal R}$}\,,& {}^{\sssh\!}\varphi^{}_a\mapsto{}^{\ssshp\!}\varphi^{}_a & \end{array} \labl{jj} which map each basis element ${}^{\sssh\!}\varphi^{}_a$ of \mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$}\ to that basis element of \mbox{$\overline{\mbox{${\cal R}$}}$}\ and \mbox{$^{\scriptscriptstyle(h')\!}{\cal R}$}\ ($h'\ge h$), respectively, which is labelled by the same weight $a\in\mbox{$_{}^{\scriptscriptstyle(h)\!\!}P$}\subseteq\mbox{$_{}^{\scriptscriptstyle(h')\!\!}P$}\subset\mbox{$\bar P$}$. For $h\preceq h'\preceq h''$, these maps satisfy \begin{equation} \jb{h'}\circ\jj h{h'} =\jb h \,, \quad \jj{h'}{h''}\circ\jj h{h'} =\jj h{h''}\, \end{equation} as well as \begin{equation} \fb{h}\circ\jb h ={\sf id}_h \,, \quad \ff{h'}{h}\circ\jj h{h'} ={\sf id}_{h} \end{equation} and \begin{equation} \fb{h}\circ\jb{h'} =\ff{h'}h \,, \quad \ff{h''}{h}\circ\jj{h'}{h''} =\ff{h'}h \,. \end{equation} We say that a generalized sequence $\psi$ in the projective limit \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}\ is {\em ultimately constant\/} iff there exists a ${h_\circ}\iN\II$ such that \begin{equation} \psi(h) = \jj{h_\circ} h\circ\psi({h_\circ}) \labl{const} (and hence for basis elements in particular ${a_h^{}}= a_{h_\circ}$) for all heights $h\ge{h_\circ}$. Now assume that $\psi_1$ and $\psi_2$ are elements of \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}\ which are ultimately constant, with associated heights ${h_{\circ,1}}$ and ${h_{\circ,2}}$, respectively. Then in particular for all heights $h$ larger than ${h_\circ}:=2\,{\rm max} ({h_{\circ,1}},{h_{\circ,2}})$ the fusion product $\psi_1(h)\star\psi_2(h)$ in \mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$}\ is isomorphic to the product $\overline\psi_1\star\overline\psi_2$ in \mbox{$\overline{\mbox{${\cal R}$}}$}, where $\overline\psi_1 :=\jb{h_\circ}\circ\psi_1({h_\circ})$, and analogously for $\overline\psi_2$. This implies that \begin{equation} (\jj{h_\circ} h\circ\psi_1({h_\circ})) \star ((\jj{h_\circ} h\circ\psi_2({h_\circ})) = \jj{h_\circ} h\circ (\psi_1({h_\circ}) \star\psi_2({h_\circ})) \labl{labl} even though $\jj{h_\circ} h$ is not a ring homomorphism, and hence $(\psi_1\star\psi_2)(h)\equiv\psi_1(h)\star\psi_2(h)=\jj{h_\circ} h\circ (\psi_1\star\psi_2)({h_\circ})$ for all $h\ge{h_\circ}$. Thus the product $\psi_1\star\psi_2$ is again ultimately constant. Also, the property of being ultimately constant is preserved upon taking (finite) linear transformations and conjugates. The set of ultimately constant elements therefore constitutes a subring of \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}. \futnot{positivity of the $\n abc$ in the subring is guaranteed once the over-all sign for the basis elements is chosen properly.} The following consideration shows that this subring is isomorphic to the fusion ring \mbox{$\overline{\mbox{${\cal R}$}}$}. First, any ultimately constant element is a linear combination of ultimately constant elements $\psi_{}^{\scriptscriptstyle(a)}$ for which $\psi_{}^{\scriptscriptstyle(a)}({h_\circ})$ is an element of the canonical basis of \mbox{$^{\scriptscriptstyle(\ho)\!}{\cal R}$}, $\psi_{}^{\scriptscriptstyle(a)}({h_\circ})={}^{\sss(\ho)\!}\varphi^{}_a$ for some $a\in\mbox{$_{}^{\scriptscriptstyle(\ho)\!\!}P$}\subset\mbox{$\bar P$}$. But there is a unique element $\psi$ of \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}\ with the latter property, because at all heights $h$ smaller than ${h_\circ}$ the value $\psi(h)$ is already fixed by imposing the requirement \erf k. Thus there is a bijective linear map between the subring of ultimately constant elements and the fusion ring \mbox{$\overline{\mbox{${\cal R}$}}$}, defined by $\bar\varphi^{}_a\mapsto\psi_{}^{\scriptscriptstyle(a)}$ for $a\!\in\!\mbox{$\bar P$}$. Moreover, the same argument which led to \erf{labl} shows that this map is in fact an isomorphism of fusion rings. As this map is provided in a canonical manner, we can actually identify the two rings. A generic element of \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}\ is {\em not\/} ultimately constant, so that the subring of ultimately constant elements is a proper subring of \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}. Thus what we have achieved is to identify the fusion ring \mbox{$\overline{\mbox{${\cal R}$}}$}\ as a proper sub-fusion ring of the projective limit ring \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}. To conclude this section, let us remark that of course we could have enlarged by hand the category ${\cal F}\!us\mbox{\scriptsize(}\mbox{$\mathfrak g$}\mbox{\scriptsize)}$\ to a larger category ${\cal F}\!u\!\!\overline{\!\!\mbox{~~}s\mbox{\scriptsize(}\mbox{$\mathfrak g$}\mbox{\scriptsize)}}$\ by just including one additional object into the category, namely the ring \mbox{$\overline{\mbox{${\cal R}$}}$}, together with the morphisms \fb h. This essentially amounts to cutting the category of rings in such a way that one is able to identify the ring \mbox{$\overline{\mbox{${\cal R}$}}$}\ as the projective limit of this category ${\cal F}\!u\!\!\overline{\!\!\mbox{~~}s\mbox{\scriptsize(}\mbox{$\mathfrak g$}\mbox{\scriptsize)}}$. We do not regard this as a viable alternative to our construction, though, since when doing so one performs manipulations which are suggested merely by one's prejudice on what the limit should look like. (Also, phenomena like level-rank dualities in fusion rings require to consider various rings for different algebra s \mbox{$\mathfrak g$}\ on the same footing; the category ${\cal F}\!u\!\!\overline{\!\!\mbox{~~}s\mbox{\scriptsize(}\mbox{$\mathfrak g$}\mbox{\scriptsize)}}$\ cannot accommodate such phenomena.) In contrast, our construction of the limit employs only the description in terms of coherent sequences, which is a natural procedure for any small category, and does not presuppose any desired features of the limit. \Sect{Representation\ theory of $\mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}$}{rep} A basic tool in the study of fusion rings is their representation\ theory. Of particular importance are the irreducible representation s, which lead in particular to the notion of (generalized) quantum dimensions. In this section we show that an analogous representation\ theory exists for the projective limit as well. In our considerations the \topo\ will again play an essential r\^ole. \subsection{One-dimensional\ representation s} Let us consider for any two $h,h'\iN\II$ with $h'=\ell h$ the injection of the label set \mbox{$_{}^{\scriptscriptstyle(h)\!\!}P$}\ (defined as in \erf{ph}) into the label set \mbox{$_{}^{\scriptscriptstyle(h')\!\!}P$}\ that is defined by multiplying the weights $a$ by a factor of $\ell$: \begin{equation} a \;\mapsto\; \ell a \end{equation} for all $a\!\in\!\mbox{$_{}^{\scriptscriptstyle(h)\!\!}P$}$. (This induces an injection ${}^{\sssh\!}\varphi^{}_a\mapsto{}^{\ssslh\!}\varphi^{}_{\ell a}$ of the distinguished basis $\mbox{$^{\scriptscriptstyle(h)}\!{\cal B}$}$ of $\mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$}$ into the distinguished basis $\mbox{$^{\scriptscriptstyle(\el h)}\!{\cal B}$}$ of $\mbox{$^{\scriptscriptstyle(\el h)\!}{\cal R}$}$. However, when this map is extended linearly to all of \mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$}, it does {\em not\/} provide a homomorphism of fusion rings.) We can use these injections to perform an {\em inductive\/} limit of the set $(\mbox{$_{}^{\scriptscriptstyle(h)\!\!}P$})_{h\in I}$ of label sets, where the set $I$ \erf i is again considered as directed via the partial ordering \erf l. We denote this inductive limit by ${}^{\scriptscriptstyle(\infty)\!\!}P$. An element $\alpha$ of ${}^{\scriptscriptstyle(\infty)\!\!}P$ can be characterized by an integrable weight ${\alpha(h)}\!\in\!\mbox{$_{}^{\scriptscriptstyle(h)\!\!}P$}$ at some suitable height $h$; at any multiple $\ell h$ of this height, the same element $\alpha$ of ${}^{\scriptscriptstyle(\infty)\!\!}P$ is then represented by the weight ${\alpha(\el h)}=\ell\,{\alpha(h)}$. In particular, quite unlike as in the case of the projective limit, each element of the inductive limit ${}^{\scriptscriptstyle(\infty)\!\!}P$ is already determined by its representative at a single height. Also note that an element $\alpha\!\in\!{}^{\scriptscriptstyle(\infty)\!\!}P$ is {\em not\/} defined at all heights $h$; in particular, for any $h\iN\II$ the set of those $\alpha\!\in\!{}^{\scriptscriptstyle(\infty)\!\!}P$ which have a representative at height $h$ is in one-to-one correspondence with the elements of \mbox{$_{}^{\scriptscriptstyle(h)\!\!}P$}, and hence is in particular finite. We will use the notation $\alpha\!\downarrow\!\!\mbox{$_{}^{\scriptscriptstyle(h)\!\!}P$}$ to indicate that $\alpha\!\in\!{}^{\scriptscriptstyle(\infty)\!\!}P$ has a representative ${\alpha(h)}\!\in\!\mbox{$_{}^{\scriptscriptstyle(h)\!\!}P$}$ at height $h$. We claim that any element of ${}^{\scriptscriptstyle(\infty)\!\!}P$ gives rise to a one-dimensional representation\ of the projective limit \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}\ of the fusion rings. To see this, we choose for a given $\alpha\!\in\!{}^{\scriptscriptstyle(\infty)\!\!}P$ a suitable height $h\iN\II$ such that $\alpha\!\downarrow\!\!\mbox{$_{}^{\scriptscriptstyle(h)\!\!}P$}$. To any coherent sequence $(\psi(l))_{l\in I}$ in the projective limit $\mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}$ we then associate the number \begin{equation} {\cal D}_\alpha(\psi) := \frac{\mbox{$^{\scriptscriptstyle(h)}\!S$}_{\psi(h),{\alpha(h)}}}{\mbox{$^{\scriptscriptstyle(h)}\!S$}_{\rho,{\alpha(h)}}} \,, \labl{cald} i.e.\ the ${\alpha(h)}$th quantum dimension of the element $\psi(h)$ of the ring \mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$}. Here we use the short-hand notation \begin{equation} \mbox{$^{\scriptscriptstyle(h)}\!S$}_{\psi(h),b}:=\sumph a \zeta_a\, \mbox{$^{\scriptscriptstyle(h)}\!S$}_{a,b} \qquad {\rm for}\;\ \psi(h)=\sumph a \zeta_a\,{}^{\sssh\!}\varphi^{}_a \end{equation} for linear combinations of $S$-matrix\ elements. Using the identities \erf D and \erf{sS} as well as ${\alpha(\el h)}=\ell{\alpha(h)}$ and the defining properties of $\psi$, we have \begin{equation} \frac{\mbox{$^{\scriptscriptstyle(\el h)}\!S$}_{\psi(\ell h),{\alpha(\el h)}}}{\mbox{$^{\scriptscriptstyle(\el h)}\!S$}_{\rho,{\alpha(\el h)}}} = \frac{(\mbox{$^{\scriptscriptstyle(\el h)}\!S$}\,D)_{\psi(\ell h),{\alpha(h)}}}{(\mbox{$^{\scriptscriptstyle(\el h)}\!S$}\,D)_{\rho,{\alpha(h)}}} = \frac{(F\,\mbox{$^{\scriptscriptstyle(h)}\!S$})_{\psi(\ell h),{\alpha(h)}}}{(F\,\mbox{$^{\scriptscriptstyle(h)}\!S$})_{\rho,{\alpha(h)}}} = \frac{\mbox{$^{\scriptscriptstyle(h)}\!S$}_{\psi(h),{\alpha(h)}}}{\mbox{$^{\scriptscriptstyle(h)}\!S$}_{\rho,{\alpha(h)}}} \,; \labl{s/s} this shows that the formula \erf{cald} yields a well-defined map from \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}\ to ${\mathbb C}$\,, i.e.\ it does not depend on the particular choice of $h$. Using the knowledge about the representation\ theory of the rings \mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$}, it then follows immediately that \begin{equation} {\cal D}_\alpha(\psi)\,{\cal D}_\alpha(\psi') = \sumph{a,b} \zeta_a^{} \zeta_b'\,\nh abc\, \frac{\mbox{$^{\scriptscriptstyle(h)}\!S$}_{c,{\alpha(h)}}}{\mbox{$^{\scriptscriptstyle(h)}\!S$}_{\rho,{\alpha(h)}}} = {\cal D}_\alpha(\psi\star\psi') \,. \end{equation} Thus the prescription \erf{cald} indeed provides us with a one-dimensional\ representation\ of $\mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}$. Let us now associate to any element $\psi$ of $\mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}$ the infinite sequence of quantum dimensions \erf{cald}, labelled by ${}^{\scriptscriptstyle(\infty)\!\!}P$; this way we obtain a map \begin{equation} {\cal D}:\quad \psi \,\mapsto\, ({\cal D}_\alpha(\psi))_{\alpha\in{}^{\scriptscriptstyle(\infty)\!\!}P} \labl{ca} from the ring \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}\ to the algebra \begin{equation} {\cal X} := \{ (\xi_\alpha)_{\alpha\in{}^{\scriptscriptstyle(\infty)\!\!}P} \!\mid\! \xi_\alpha\!\in\!{\mathbb C}\, \} \end{equation} of all countably infinite sequences of complex numbers. Since we are now dealing with complex numbers rather than only integers, it is natural to consider instead of the fusion ring \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}\ the corresponding algebra over ${\mathbb C}$\,, to which we refer as the {\em fusion algebra\/} \mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}. (For simplicity we regard \mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}\ as an algebra over ${\mathbb C}$\,. In principle it would be sufficient to consider it over a certain subfield of ${\mathbb C}$\ generated by appropriate roots of unity.) It is then evident that the map ${\cal D}\!:\,\mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}\to{\cal X}$ defined by \erf{ca} is an algebra homomorphism. (We continue to use the symbol ${\cal D}$. More generally, below we will always assume that the various maps to be used, such as the projection \erf3, are continued ${\mathbb C}$\,-linearly from the fusion rings \mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$}\ to the associated fusion algebras \mbox{$^{\scriptscriptstyle(h)}\!\!{\cal A}$}, and use the same symbols for these extended maps as for the original ones.) \subsection{An isomorphism between \mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}\ and ${\cal X}$} In this subsection we show that the map ${\cal D}$ introduced above even constitutes an {\em iso\/}morphism between the complex algebras \mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}\ and ${\cal X}$: \begin{equation} {\cal D}:\quad \mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$} \;\stackrel{\cong}{\longrightarrow}\; {\cal X} \,. \labl{dax} Injectivity of ${\cal D}$ is easy to check. Suppose we have ${\cal D}(\psi)=0$. Fix any $h\!\in\! I$; then all quantum dimensions of the element $\psi(h)$ of \mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$}\ vanish. From the properties of the fusion ring \mbox{$^{\scriptscriptstyle(h)\!}{\cal R}$}\ it then follows immediately that $\psi(h)=0$. This is true for all $h\!\in\! I$, and hence we have $\psi=0$. This proves injectivity. To show also surjectivity requires more work. We first need to introduce the elements \begin{equation} {\rm e}_a \equiv {}^{\scriptscriptstyle(h)\!}{\rm e}_a := \mbox{$^{\scriptscriptstyle(h)}\!S$}_{\rho,a} \sumph b \mbox{$^{\scriptscriptstyle(h)}\!S$}^*_{a,b}\, {}^{\sssh\!}\varphi^{}_b \;\in \mbox{$^{\scriptscriptstyle(h)}\!\!{\cal A}$} \labl{ea} of the fusion algebras at height $h$. These elements are idempotents, i.e.\ obey \begin{equation} {\rm e}_a\star {\rm e}_b = \delta_{a,b}\, {\rm e}_{a} \,. \labl{eee} Owing to the unitarity of the modular transformation matrix $S$, the idempotents $\{{}^{\scriptscriptstyle(h)\!}{\rm e}_a\!\mid\! a\!\in\!\mbox{$_{}^{\scriptscriptstyle(h)\!\!}P$}\}$ form a basis of the fusion algebra\ \mbox{$^{\scriptscriptstyle(h)}\!\!{\cal A}$}, and they constitute a partition of the unit element, in the sense that \begin{equation} \sumph a {}^{\scriptscriptstyle(h)\!}{\rm e}_a = {}^{\sssh\!}\varphi^{}_\rho \,. \labl{pou} Also, for any element $\psi\!\in\!\mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}$ with $\psi(h)={\rm e}_a$ and any $\alpha\!\in\!{}^{\scriptscriptstyle(\infty)\!\!}P$ with $\alpha\!\downarrow\!\!\mbox{$_{}^{\scriptscriptstyle(h)\!\!}P$}$ we have \begin{equation} {\cal D}_\alpha(\psi) = \delta_{a,{\alpha(h)}} \,. \labl{pea} We now study how the idempotents ${\rm e}_{{\alpha(h)}}$ behave under the projection \erf3. First, when $\alpha\!\in\!{}^{\scriptscriptstyle(\infty)\!\!}P$ has a representative ${\alpha(h)}$ at height $h$, then for every positive integer $\ell$ we have, using the first of the identities \erf{sS}, \begin{equation} \begin{array}{ll} \ff{\el h}h({\rm e}_{{\alpha(\el h)}}) \!\! &= {\mbox{$^{\scriptscriptstyle(\el h)}\!S$}_{\rho,{\alpha(\el h)}}} \sumplh A \!\!\mbox{$^{\scriptscriptstyle(\el h)}\!S$}^*_{{\alpha(\el h)},A}\ff{\el h}h(\phi^{}_A) = {\mbox{$^{\scriptscriptstyle(\el h)}\!S$}_{\rho,\ell{\alpha(h)}}} \sumplh A\, \sumph b \! \mbox{$^{\scriptscriptstyle(\el h)}\!S$}^*_{{\alpha(\el h)},A} \P Ab\, \varphi^{}_b \\{}\\[-.3em] &= \ell^{-r/2}\cdot {\mbox{$^{\scriptscriptstyle(h)}\!S$}_{\rho,{\alpha(h)}}} \sumph b\!(\mbox{$^{\scriptscriptstyle(\el h)}\!S$}^* F)_{{\alpha(\el h)},b} \, \varphi^{}_b = {\mbox{$^{\scriptscriptstyle(h)}\!S$}_{\rho,{\alpha(h)}}} \sumph{b,c} \! D_{\ell{\alpha(h)},c}\mbox{$^{\scriptscriptstyle(h)}\!S$}^*_{c,b}\, \varphi^{}_b \\{}\\[-.7em] &= \mbox{$^{\scriptscriptstyle(h)}\!S$}_{\rho,{\alpha(h)}} \sumph b \mbox{$^{\scriptscriptstyle(h)}\!S$}^*_{{\alpha(h)},b}\, \varphi^{}_b = {\rm e}_{{\alpha(h)}} \, . \end{array}\labl{ee} On the other hand, when $\alpha$ has a representative at height $h$, but not at height $h'$, we can compute as follows. Since $hh'$ is a multiple of $h$, $\alpha$ has a representative ${\alpha(hh')}$ at height $hh'$. Thus we can repeat the previous calculation to deduce that \begin{equation} \begin{array}{ll} \ff{hh'}{h'}({\rm e}_{{\alpha(hh')}}) \!\!& = \mbox{$^{\scriptscriptstyle(hh')}\!S$}_{\rho,{\alpha(hh')}} \cdot h^{r/2} \sumphp b (D\,\mbox{$^{\scriptscriptstyle(h')}\!S$})_{{\alpha(hh')},b} \, {}^{\ssshp\!}\varphi^{}_b \\{}\\[-.7em] & = (h/h')^{r/2}\cdot \mbox{$^{\scriptscriptstyle(h)}\!S$}_{\rho,{\alpha(h)}} \sumphp{b,c} \delta_{h'{\alpha(h)},hc} \mbox{$^{\scriptscriptstyle(h')}\!S$}_{c,b}\, {}^{\ssshp\!}\varphi^{}_b \,. \end{array} \end{equation} Now in the sum over $c$ on the right hand side\ one has a contribution only if $c=h'{\alpha(h)}/h={\alpha(hh')}/h$ is an element of the label set \mbox{$_{}^{\scriptscriptstyle(h')\!\!}P$}\ at height $h'$. But in this case we would conclude that $\alpha$ has in fact a representative at height $h'$, namely ${\alpha(h')}=c$, which contradicts our assumption. Therefore we conclude that in the case under consideration we have $\ff{hh'}{h'}({\rm e}_{{\alpha(hh')}})=0.$ Together with the result \erf{ee} it follows that by setting \begin{equation} {\rm e}_\alpha(h):= \left\{ \begin{array}{ll} 0 & \mbox{if $\alpha\!\in\!{}^{\scriptscriptstyle(\infty)\!\!}P$ has no representative at height $h$}\,, \\[.1em] {\rm e}_{{\alpha(h)}} & {\rm else}\,, \end{array} \right. \labl{eal} we obtain an element ${\rm e}_\alpha$ of the projective limit \mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}. Moreover, according to the relation \erf{pea} the map \erf{dax} acts on ${\rm e}_\alpha\!\in\!\mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}$ as \begin{equation} \mbox{\large(} {\cal D}({\rm e}_\alpha) \mbox{\large)}_\beta = \delta_{\alpha,\beta} \labl{peb} for all $\alpha,\beta\!\in\!{}^{\scriptscriptstyle(\infty)\!\!}P$, and the ${\rm e}_\alpha$ provide a partition of the unit element, analogously as in \erf{pou}, \begin{equation} \sum_{\alpha\in{}^{\scriptscriptstyle(\infty)\!\!}P} {\rm e}_\alpha = \psi_\circ \,; \end{equation} here the sum is to be understood as a limit of finite sums in the \topo. Now for each $h\iN\II$ let us define the map \,$g_h\!:\, {\cal X}\to\mbox{$^{\scriptscriptstyle(h)}\!\!{\cal A}$}$ by $(\xi_\alpha)_{\alpha\in{}^{\scriptscriptstyle(\infty)\!\!}P}\mapsto \sum_{\scriptstyle\beta\in{}^{\scriptscriptstyle(\infty)\!\!}P \atop \scriptstyle\beta\downarrow_{}^{\scriptscriptstyle(h)\!}P} \xi_\beta\, {\rm e}_{{\beta(h)}}$. Since $\beta\!\downarrow\!\!\mbox{$_{}^{\scriptscriptstyle(\el h)\!\!}P$}$ \,if\, $\beta\!\downarrow\!\!\mbox{$_{}^{\scriptscriptstyle(h)\!\!}P$}$, we then have \begin{equation} \ff{\el h}h \circ g_{\ell h}^{}((\xi_\alpha)) = \sum_{\scriptstyle\beta\in{}^{\scriptscriptstyle(\infty)\!\!}P \atop \scriptstyle\beta\downarrow_{}^{\scriptscriptstyle(\el h)\!}P}\xi_\beta\, {\rm e}_{\beta}(h) = \sum_{\scriptstyle\beta\in{}^{\scriptscriptstyle(\infty)\!\!}P \atop \scriptstyle\beta\downarrow_{}^{\scriptscriptstyle(h)\!}P}\xi_\beta\, {\rm e}_{{\beta(h)}} = g_h((\xi_\alpha)) \end{equation} for all positive integers $\ell$. Analogously we can define a map \begin{equation} g:\quad {\cal X} \to \mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}\,,\qquad (\xi_\alpha)_{\alpha\in{}^{\scriptscriptstyle(\infty)\!\!}P}^{}\mapsto \!\!\sum^{}_{\beta\in{}^{\scriptscriptstyle(\infty)\!\!}P} \xi_\beta\,{\rm e}_\beta \end{equation} with similar properties. As a consequence of the relation \erf{peb} one finds that this map satisfies \begin{equation} {\cal D} \circ g = {\sf id}_{{\cal X}}^{} \,. \end{equation} This implies that the injective map ${\cal D}$ is also surjective (and that $g$ is injective). Thus we have proven the isomorphism \erf{dax}. \subsection{Semi-simplicity} It is known \cite{kawA} that the fusion algebra s \mbox{$^{\scriptscriptstyle(h)}\!\!{\cal A}$}\ at finite heights $h$ are semi-simple\ associative algebra s. In this subsection we show that in a suitable topological sense the same statement holds for the projective limit \mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}, too. We first combine the identity \erf{eee} and the definition \erf{eal} of the element ${\rm e}_\alpha$ of \mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}\ with the fact that the idempotents ${\rm e}_a$ form a basis of \mbox{$^{\scriptscriptstyle(h)}\!\!{\cal A}$}. This way we learn that for all $\psi\!\in\! \mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}$ and all heights $h\iN\II$ the fusion product $({\rm e}_\alpha\star\psi)(h)= {\rm e}_\alpha(h)\star\psi(h)$ is proportional to ${\rm e}_\alpha(h)$. Thus for each $\alpha\!\in\!{}^{\scriptscriptstyle(\infty)\!\!}P$ the span \begin{equation} {\cal I}_\alpha := \langle{\rm e}_\alpha\rangle \labl{Id} of $\{{\rm e}_\alpha\}$ is a one-dimensional\ twosided ideal of the projective limit, i.e.\ we have $\mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}\,{\cal I}_\alpha={\cal I}_\alpha\mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}\subseteq{\cal I}_\alpha$. We claim that when we endow the algebra\ with the \topo, then in fact \mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}\ is the closure of the direct sum of the ideals \erf{Id} in this topology: \begin{equation} \mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$} = \overline{\bigoplus_{\alpha\in{}^{\scriptscriptstyle(\infty)\!\!}P} {\cal I}_\alpha} \,. \labl{clos} (In particular, the idempotents ${\rm e}_\alpha$ form a topological basis of \mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}.) To prove this, we first recall from subsection \ref{topo} that in the \topo\ each open set in \mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}\ is a union of elements of the set $\mbox{$\Omega$}=\{ \efm h(M) \!\mid\! h\!\in\!\mbox{$I$},\; M\;{\rm open\;in}\;\mbox{$^{\scriptscriptstyle(h)}\!\!{\cal A}$}\}$ of all pre-images of all open sets in any of the fusion algebra s \mbox{$^{\scriptscriptstyle(h)}\!\!{\cal A}$}. Here we assume that we have already chosen a topology on each of the fusion algebra s \mbox{$^{\scriptscriptstyle(h)}\!\!{\cal A}$}. (Actually the choice of this topology on \mbox{$^{\scriptscriptstyle(h)}\!\!{\cal A}$}\ will not be important; for definiteness, we may take the discrete one, as in the case of fusion rings, or also the metric topology of \mbox{$^{\scriptscriptstyle(h)}\!\!{\cal A}$}\ as a finite-dimensional\ complex vector space.) Consider now an arbitrary element $\xi\!\in\!\mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}$, which because of the isomorphism $\mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}\cong{\cal X}$ we can write as $\xi=(\xi_\alpha)_{\alpha\in {}^{\scriptscriptstyle(\infty)\!\!}P}$. Adopting some definite numbering ${}^{\scriptscriptstyle(\infty)\!\!}P=\{{\alpha_m}\!\mid\! m\!\in\!{\dl N}\}$ of the countable set ${}^{\scriptscriptstyle(\infty)\!\!}P$, for $n\!\in\!{\dl N}$ we define \begin{equation} \hat\xi_n := \sum_{m\le n} \xi_{\alpha_m} {\rm e}_\alpham \ \in\; \bigoplus_{\alpha\in {}^{\scriptscriptstyle(\infty)\!\!}P} {\cal I}_\alpha \,. \end{equation} To prove our assertion, we must then show that for every $h\iN\II$ and every open set $M\subseteq\mbox{$^{\scriptscriptstyle(h)}\!\!{\cal A}$}$ which satisfy $f_h(\xi)\!\in\! M$ we have $\hat\xi_n\!\in\! f_h^{-1}(M)$, i.e.\ \begin{equation} f_h(\hat\xi_n) \in M \,, \end{equation} for all but finitely many $n$. Now by direct calculation we obtain \begin{equation} f_h(\hat\xi_n) = \sum_{\scriptstyle m\le n \atop\ \scriptstyle{\alpha_m}\downarrow_{}^{\scriptscriptstyle(h)\!}P} \!\! \xi_{\alpha_m} {}^{\scriptscriptstyle(h)\!}{\rm e}_{\alpham(h)} \,; \end{equation} this is a finite sum, and for sufficiently large $n$ it becomes independent of $n$ because only finitely many $\alpha\!\in\!{}^{\scriptscriptstyle(\infty)\!\!}P$ have a representative in \mbox{$_{}^{\scriptscriptstyle(h)\!\!}P$}. In fact, for sufficiently large $n$ we simply have \begin{equation} f_h(\hat\xi_n) = \sum_{\alpha\downarrow_{}^{\scriptscriptstyle(h)\!}P} \xi_\alpha\, {}^{\scriptscriptstyle(h)\!}{\rm e}_{\alpha(h)} \equiv f_h(\xi) \,. \end{equation} Since $f_h(\xi)\!\in\! M$, this immediately shows that indeed $f_h(\hat\xi_n) \!\in\! M$ for almost all $n$, and hence the proof is completed. (Note that the fact that $f_h(\hat\xi_n)$ ultimately becomes equal to $f_h(\xi)$ holds for any chosen topology of \mbox{$^{\scriptscriptstyle(h)}\!\!{\cal A}$}, and hence the conclusion is indeed independent of that topology.) \subsection{Simple and semi-simple\ modules}\label{Sss} The representation\ theory of \mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}\ can now be developed by following the same steps as in the representation\ theory of semi-simple\ algebras. However, when considering modules $V$ over \mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}, it is natural to restrict one's attention from the outset to {\em continuous\/} modules, i.e.\ to modules which are topological vector spaces and on which the representation\ of \mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}\ is continuous (in particular, every element of \mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}\ is represented by a continuous map). We will do so, and suppress the qualification `continuous' from now on. The one-dimensional\ ideals ${\cal I}_\alpha$ are simple modules over \mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}\ under the (left or right) regular representation. Our first result is now that these one-dimensional\ modules already provide us with all simple modules, i.e.\ that every simple \mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}-module $L$ satisfies \begin{equation} L \,\cong\,{\cal I}_\alpha \end{equation} for some $\alpha\!\in\!{}^{\scriptscriptstyle(\infty)\!\!}P$. To show this, we first observe that if $L\not\cong{\cal I}_\alpha$, then ${\cal I}_\alpha L = 0$. Namely, since ${\cal I}_\alpha$ is an ideal of \mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}, we have $\mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}\,{\cal I}_\alpha L\subseteq{\cal I}_\alpha L$; thus ${\cal I}_\alpha L$ is a submodule of $L$, which by the simplicity of $L$ implies that either ${\cal I}_\alpha L = L$ or ${\cal I}_\alpha L=0$. In the former case, ${\cal I}_\alpha L=L$, we can find a vector $y\!\in\! L$ such that the space ${\cal I}_\alpha y$ is not zero-dimensional. Indeed, because of $\mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}\,{\cal I}_\alpha\,y\subseteq {\cal I}_\alpha\,y\subseteq L$ this space is a submodule of $L$, and hence by the simplicity of $L$ it must be equal to $L$. It follows that the map from ${\cal I}_\alpha$ to $L$ defined by $\lambda \mapsto \lambda y$ is surjective. Since $L$ is simple, by Schur's lemma this implies that it is even an isomorphism. This shows that $L\cong{\cal I}_\alpha$ when ${\cal I}_\alpha L=L$, and hence ${\cal I}_\alpha L=0$ when $L\not\cong{\cal I}_\alpha$.\\ Suppose now that $L$ is a non-zero simple module and is not isomorphic to any ${\cal I}_\alpha$. Then $\bigoplus_\alpha{\cal I}_\alpha L = 0$; since $L$ is a continuous module, we can take the closure of this relation, so as to find that $\mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}\,L =0$. But we have $L\subseteq\mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}\,L$, and hence this would imply that $L=0$, which is a contradiction. Hence we learn that indeed, up to isomorphism, the ideals ${\cal I}_\alpha$ of \mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}\ exhaust all the simple modules over \mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}. Next we consider modules $V$ over \mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}\ which can be obtained from families of simple modules. Similarly as in \cite[\S XVII.2]{LAng} one can show that the following conditions are equivalent: \begin{quote} $(i)$~~~~$V$ is the closure of the sum of a family of simple submodules.\\[.3em] $(ii)$~~~$V$ is the closure of the direct sum of a family of simple submodules. \\[.3em] $(iii)$\hsp{2.85}Every closed submodule $W$ of $V$ is a direct summand, i.e.\ there exists\\\hsp{10.48}a closed submodule $W'$ such that $V = W \oplus W'$. \end{quote} Any (continuous) module fulfilling these equivalent conditions will be referred to as a {\emsemi-simple\/} module. The equivalence of $(i)$\,$\mbox{-\hspace{-.66 mm}-}$\,$(iii)$ is proven as follows. First, if $V=\overline{\sum_{i\in J}L_i}$ is the closure of a (not necessarily direct) sum of simple submodules $L_i$, denote by $J'$ a maximal subset of $J$ such that $V':=\sum_{j\in J'}L_j$ is a direct sum. Since the intersection of $V'$ with any of the simple modules $L_i$ is a submodule of $L_i$, the maximality of $J'$ implies that $i\!\in\! J'$ and hence in fact $J'=J$. Thus $(i)$ implies $(ii)$. \\ Second, if $W$ is a submodule of $V$, let $J''$ be the maximal subset of $J$ such that the sum $W+\sum_{j\in J''}L_j$ is direct. Then the same arguments as before show that $V = \overline{W \oplus \bigoplus_{j\in J''}L_j}$. If, furthermore, $W$ is closed, then it follows that $V = \overline{W}\oplus \overline{\bigoplus_{j\in J''}L_j} = W\oplus\overline{\bigoplus_{j\in J''}L_j}$. This shows that$(ii)$ implies $(iii)$.\,% \futnote{It is indeed necessary to require $W$ to be closed. Consider e.g.\ the case $V=\mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}$ and $W=\bigoplus_\alpha{\cal I}_\alpha$. The submodule $W$ is neither closed nor does it have a complement.} \noindent Third, assume that $V$ is a non-zero module which satisfies $(iii)$, and let $v$ be a non-zero vector in $V$. The kernel of the homomorphism $\mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}\to\mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}\, v$ is a closed ideal of \mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}, which in turn is contained in a maximal closed ideal ${\cal J}\!\subset\!\mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}$ that is strictly contained in \mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}. One then has $V={\cal J}v \oplus W$ and $\mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}\,v={\cal J}v \oplus (W\cap\mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}\, v)$ with some submodule $W\subset V$. Now $W\cap\mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}\, v$ is simple because ${\cal J}v$ is maximal in $\mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}\, v$; thus $V$ contains a simple submodule. Next let $V' \ne 0 $ be the submodule of $V$ that is the closure of the sum of all simple submodules of $V$. If $V'$ were not all of $V$, then one would have $V=V'\oplus V''$ with $V''\ne0$; but by the same reasoning as before, $V''$ then would contain a simple submodule, in contradiction to the definition of $V'$. Thus $V'=V$, so we see that $(iii)$ implies $(i)$. \subsection{Arbitrary modules} With the characterization of semi-simple\ modules above, we are now in a position to study arbitrary modules of \mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}, in an analogous manner as in \cite[\S XVII.4]{LAng}. Let us first assume that $W$ is a closed submodule of a semi-simple\ module $V$, and denote by $W'$ the closure of the direct sum of all simple submodules of $W$. Then there is a submodule $V'$ of $V$ such that $V= W'\oplus V'$. Every $w\!\in\! W$ can be uniquely written as $w=w'+v'$ with $w'\!\in\! W'$ and $v'\!\in\! V'$. Because of $v'=w-w'\!\in\! W$ we thus have $W=W'\oplus(W\cap V')$. The module $W\cap V'$ is a closed submodule of $W$. If it were non-zero, it would therefore (by the same reasoning as in the proof of `$(iii)\to(i)$' in subsection \ref{Sss}) contain a simple submodule, in contradiction with the definition of $W$. Thus we learn that $W=W'$, or in other words: \begin{quote} Every closed submodule of a semi-simple\ \mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}-module is semi-simple. \end{quote} \noindent Next we consider again a closed submodule $W$ of a semi-simple\ module $V$, and investigate the quotient module $V/W$. There is a closed submodule $W'$ such that $V$ is the direct sum $V=W\oplus W'$. Now the projection $V \to V/W$ induces a continuous isomorphism from $W'$ to $V/W$. Furthermore, according to the result just obtained, $W'$ is semi-simple. Thus we have shown: \begin{quote} Every quotient module of a semi-simple\ \mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}-module with respect to \\ a closed submodule is semi-simple. \end{quote} \noindent Now any arbitrary \mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}-module can be regarded as a quotient module of a suitable free module modulo a closed submodule. Moreover, every free \mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}-module is the closure of a direct sum of countably many copies of \mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}\ and hence is a semi-simple\ module. The two previous results therefore imply: \begin{quote} Every \mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}-module is semi-simple. \end{quote} Finally we consider again an arbitrary \mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}-module $V$. We denote by $V_\alpha$ the closure of the direct sum of all those submodules of $V$ which are isomorphic to the simple \mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}-module ${\cal I}_\alpha$. Since each simple module over \mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}\ is isomorphic to some ${\cal I}_\alpha$, any simple submodule of $V$ is contained in $V_\beta$ for some $\beta\!\in\!{}^{\scriptscriptstyle(\infty)\!\!}P$. Now every \mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}-module is semi-simple\ and hence the closure of the direct sum of its simple submodules. Thus we learn that \begin{equation} V = \overline{ \bigoplus_{\alpha\in{}^{\scriptscriptstyle(\infty)\!\!}P} V_\alpha } \, . \end{equation} Moreover, we have ${\rm e}_\beta V_\alpha=\delta_{\alpha,\beta}V_\alpha$ for all $\alpha,\beta\!\in\!{}^{\scriptscriptstyle(\infty)\!\!}P$, and hence \begin{equation} V_\alpha = {\rm e}_\alpha V = {\cal I}_\alpha V \, . \end{equation} As a consequence, we see that: \begin{quote} Every \mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}-module $V$ can be written as \begin{equation} V = \overline{\bigoplus_{\alpha\in{}^{\scriptscriptstyle(\infty)\!\!}P} {\cal I}_\alpha V} = \overline{\bigoplus_{\alpha\in{}^{\scriptscriptstyle(\infty)\!\!}P} {\rm e}_\alpha V} \,, \end{equation} and for each $\alpha\!\in\!{}^{\scriptscriptstyle(\infty)\!\!}P$ the submodule ${\cal I}_\alpha V$ is the closure of the direct sum of all submodules of $V$ that are isomorphic to ${\cal I}_\alpha$. \end{quote} \noindent We can conclude that the structure of any arbitrary (continuous) module over \mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}\ is known explicitly, i.e.\ we have developed the full (topological) representation\ theory of \mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}. \subsection{Diagonalization} {}From the definition \erf{eal} of ${\rm e}_\alpha\!\in\!\mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}$ and the basic property \erf{eee} of the idempotents ${\rm e}_a\!\in\!\mbox{$^{\scriptscriptstyle(h)}\!\!{\cal A}$}$ it follows that the elements ${\rm e}_\alpha$ of \mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}\ are again idempotents: \begin{equation} {\rm e}_\alpha \star {\rm e}_\beta = \delta_{\alpha,\beta}\,{\rm e}_\alpha \end{equation} for all $\alpha,\beta\!\in\!{}^{\scriptscriptstyle(\infty)\!\!}P$. In other words, by the basis transformation from the distinguished basis \mbox{$^{\scriptscriptstyle(\infty)\!\!}{\cal B}$}\ of the fusion algebra\ \mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}\ to the basis of idempotents one diagonalizes the fusion rules of \mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}, precisely as in the case of the algebra s \mbox{$^{\scriptscriptstyle(h)}\!\!{\cal A}$}\ at finite level. Indeed, by combining the definitions \erf{ea} and \erf{eal} we can describe the transformation from \mbox{$^{\scriptscriptstyle(\infty)\!\!}{\cal B}$}\ to the basis of idempotents ${\rm e}_\alpha$ explicitly. Namely, for any $\psi=(\psi(h))_{h\inI}\!\in\!\mbox{$^{\scriptscriptstyle(\infty)\!\!}{\cal B}$}$ we have \begin{equation} \psi = \sum_{\alpha\in{}^{\scriptscriptstyle(\infty)\!\!}P}\! {}^{\sss(\infty)\!}Q_{\psi,\alpha} \, {\rm e}_\alpha \,, \end{equation} with \begin{equation} {}^{\sss(\infty)\!}Q_{\psi,\alpha}:= \frac{\mbox{$^{\scriptscriptstyle(h)}\!S$}_{\psi(h),{\alpha(h)}}}{\mbox{$^{\scriptscriptstyle(h)}\!S$}_{\rho,{\alpha(h)}}} \,, \end{equation} where $h\iN\II$ is a height at which $\alpha$ has a representative. Note that owing to the relation \erf{s/s} the quotient ${}^{\sss(\infty)\!}Q_{\psi,\alpha}$ does not depend on the particular choice of $h$. (This just rephrases the fact that the map $\cal D$ is an isomorphism.) For any $\psi,\psi'\!\in\!\mbox{$^{\scriptscriptstyle(\infty)\!\!}{\cal B}$}$ we thus have \begin{equation} \psi\star\psi'= \sum_{\alpha\in{}^{\scriptscriptstyle(\infty)\!\!}P} \frac{\mbox{$^{\scriptscriptstyle(h)}\!S$}_{\psi(h),{\alpha(h)}}} {\mbox{$^{\scriptscriptstyle(h)}\!S$}_{\rho,{\alpha(h)}}} \frac{\mbox{$^{\scriptscriptstyle(h)}\!S$}_{\psi'(h),{\alpha(h)}}}{\mbox{$^{\scriptscriptstyle(h)}\!S$}_{\rho,{\alpha(h)}}} \, {\rm e}_\alpha = \sum_{\chi\in {}^{\scriptscriptstyle(\infty)\!}{\cal B}} \mbox{\Large(} \!\!\sum_{\alpha\in{}^{\scriptscriptstyle(\infty)\!\!}P} {}^{\sss(\infty)\!}Q_{\psi,\alpha} {}^{\sss(\infty)\!}Q_{\psi',\alpha} \, {}^{\sss(\infty)\!}Q^-_{\alpha,\chi}\, \mbox{\Large)} \, \chi \,, \end{equation} where ${}^{\sss(\infty)\!}Q^-$ is the matrix for the inverse basis transformation, \begin{equation} {\rm e}_\alpha = \sum_{\psi\in {}^{\scriptscriptstyle(\infty)\!}{\cal B}} {}^{\sss(\infty)\!}Q^-_{\alpha,\psi}\, \psi \,. \end{equation} In other words, the fusion rule coefficient s of the projective limit \mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}\ can be written as \begin{equation} _{}^{\scriptscriptstyle(\infty)}\!\!\n\psi{\psi'}{\psi''} = \sum_{\alpha\in{}^{\scriptscriptstyle(\infty)\!\!}P} {}^{\sss(\infty)\!}Q_{\psi,\alpha}\, {}^{\sss(\infty)\!}Q_{\psi',\alpha}\, {}^{\sss(\infty)\!}Q^-_{\alpha,\psi''} \, . \end{equation} This is nothing but the analogue of the Verlinde formula \cite{verl2} that is valid for the fusion rule coefficient s of the fusion algebra s \mbox{$^{\scriptscriptstyle(h)}\!\!{\cal A}$}. Note that already at finite height $h$ the two indices which label the rows and columns, respectively, of the matrix $\mbox{$^{\scriptscriptstyle(h)}\!S$}$ which diagonalizes the fusion rules are a priori of a rather different nature. Namely, one of them labels the elements of the distinguished basis \mbox{$^{\scriptscriptstyle(h)}\!{\cal B}$}, while the other labels the inequivalent one-dimensional\ irreducible representation s of \mbox{$^{\scriptscriptstyle(h)}\!\!{\cal A}$}. It is a quite non-trivial \futnot{and not well understood} property of the fusion algebra s which arise in rational conformal field theory\ (and is a prerequisite for the modularity of those fusion algebra s) that nevertheless the diagonalizing matrix can be chosen such that it is symmetric, so that in particular the two kinds of labels can be treated on an equal footing \cite{jf24}. Our results clearly display that this nice feature of the finite height fusion algebra s \mbox{$^{\scriptscriptstyle(h)}\!\!{\cal A}$}\ is not shared by their non-rational limit \mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}; in the case of \mbox{$_{}^{\scriptscriptstyle(\infty)\!\!\!}{\cal A}$}, there seems to be no possibility to identify the two sets \mbox{$^{\scriptscriptstyle(\infty)\!\!}{\cal B}$}\ and ${}^{\scriptscriptstyle(\infty)\!\!}P$ which label the elements of the distinguished basis and the one-dimensional\ irreducible representation s, respectively, with each other. On the other hand, our results show that the projective limit \mbox{$^{\scriptscriptstyle(\infty)\!}{\cal R}$}\ that we constructed in this paper still possesses all those structural properties of a modular fusion ring which can reasonably be expected to survive in the limit of infinite level. \def\,\linebreak[0]} \def\wB {$\,$\wb{\,\linebreak[0]} \def\wB {$\,$\,\linebreak[0]} \def\wB {$\,$\wb} \newcommand\Bi[1] {\bibitem{#1}} \newcommand\Erra[3] {\,[{\em ibid.}\ {#1} ({#2}) {#3}, {\em Erratum}]} \newcommand\BOOK[4] {{\em #1\/} ({#2}, {#3} {#4})} \newcommand\J[5] {\ {\sl #5}, {#1} {#2} ({#3}) {#4} } \newcommand\Prep[2] {{\sl #2}, preprint {#1}} \defJ.\ Fuchs {J.\ Fuchs} \defFortschr.\wb Phys. {Fortschr.\,\linebreak[0]} \def\wB {$\,$\wb Phys.} \defHelv.\wb Phys.\wB Acta {Helv.\,\linebreak[0]} \def\wB {$\,$\wb Phys.\wB Acta} \newcommand\npbF[5] {{\sl #5}, Nucl.\wb Phys.\ B\ {#1} [FS{#2}] ({#3}) {#4}} \defNucl.\wb Phys.\ B (Proc.\wb Suppl.) {Nucl.\,\linebreak[0]} \def\wB {$\,$\wb Phys.\ B (Proc.\,\linebreak[0]} \def\wB {$\,$\wb Suppl.)} \defNuovo\wB Cim. {Nuovo\wB Cim.} \defNucl.\wb Phys.\ B {Nucl.\,\linebreak[0]} \def\wB {$\,$\wb Phys.\ B} \defPhys.\wb Lett.\ B {Phys.\,\linebreak[0]} \def\wB {$\,$\wb Lett.\ B} \defJ.\wb Math.\wb Phys. {J.\,\linebreak[0]} \def\wB {$\,$\wb Math.\,\linebreak[0]} \def\wB {$\,$\wb Phys.} \defCom\-mun.\wb Math.\wb Phys. {Com\-mun.\,\linebreak[0]} \def\wB {$\,$\wb Math.\,\linebreak[0]} \def\wB {$\,$\wb Phys.} \defAlgebra {Algebra} \def{Academic Press} {{Academic Press}} \def{Addi\-son\hy Wes\-ley} {{Addi\-son$\mbox{-\hspace{-.66 mm}-}$ Wes\-ley}} \defalgebra {algebra} \def{Berlin} {{Berlin}} \def{Birk\-h\"au\-ser} {{Birk\-h\"au\-ser}} \def{Cambridge} {{Cambridge}} \def{Cambridge University Press} {{Cambridge University Press}} \deffusion rule {fusion rule} \def{Gordon and Breach} {{Gordon and Breach}} \newcommand{\inBO}[7] {in:\ {\em #1}, {#2}\ ({#3}, {#4} {#5}), p.\ {#6}} \defInfinite-dimensional {Infinite-dimensional} \def{Kluwer Academic Publishers} {{Kluwer Academic Publishers}} \def{New York} {{New York}} \defQuantum\ {Quantum\ } \defquantum group {quantum group} \def{Sprin\-ger Verlag} {{Sprin\-ger Verlag}} \defsym\-me\-tries {sym\-me\-tries} \deftransformation {transformation} \defWZW\ {WZW\ } \vskip5em \small \noindent{\bf Acknowledgement.} \ We are grateful to I.\ Kausz and B.\ Pareigis for helpful comments. \newpage
proofpile-arXiv_065-433
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} This paper is devoted to the spectral decomposition of the space of sections $L^2(Y,V_Y(\gamma))$ of a locally homogenous bundle $V_Y(\gamma)$ over a locally symmetric space $Y$ of rank one of infinite volume with respect to locally invariant differential operators. Let $G$ be a real semisimple linear connected Lie group of real rank one with finite center. Let $K\subset G$ be a maximal compact subgroup. Then $X:=G/K$ is a Riemannian symmetric space of negative curvature. Let $\Gamma\subset G$ be a convex-cocompact, torsion-free, discrete subgroup and $Y:=\Gamma\backslash X$ be the corresponding locally symmetric space. Let $(\gamma,V_\gamma)$ be a finite-dimensional unitary representation of $K$. Then we form the (locally) homogeneous vector bundle $V(\gamma):=G\times_KV_\gamma$ ($V_Y(\gamma):=\Gamma\backslash G\times_KV_\gamma$). Let ${\bf g}$ denote the Lie algebra of $G$, ${\cal U}({\bf g})$ the universal enveloping algebra of ${\bf g}$ and ${\cal Z}$ its center. Through the left regular action of ${\cal U}({\bf g})$ on $C^\infty(X,V(\gamma))$ any $A\in {\cal Z}$ gives rise to an $G$-invariant differential operator $A_\gamma$. This operator descends to $C^\infty(Y,V_Y(\gamma))$. Let ${\cal Z}_\gamma$ denote the algebra $\{A_\gamma\:|\: A\in {\cal Z}\}$. By $\nabla$ we denote the canonical invariant connection of $V(\gamma)$. The algebra ${\cal Z}_\gamma$ is a finite extension of the algebra ${\bf C}(\Delta)$, where $\Delta=\nabla^*\nabla$ is the Bochner Laplace operator on $V(\gamma)$. We employ a suitable invariant scalar product on ${\bf g}$ in order to normalize the Riemannian metric of $X$ and the Casimir operators $\Omega_G$ and $\Omega_K$ of $G$ and $K$. The Casimir operators are related with the Laplacian by $\Delta=-\Omega_G+\gamma(\Omega_K)$. We view ${\cal Z}_\gamma$ as an algebra of unbounded operators on the Hilbert spaces $L^2(X,V(\gamma))$ (resp. $L^2(Y,V_Y(\gamma))$) with common domain $C_c^\infty(X,V(\gamma))$ (resp. $C_c^\infty(Y,V_Y(\gamma))$). Since $X$ (resp. $Y$) are complete any formally selfadjoint locally invariant elliptic operator is essentially selfadjoint on the domain $C^\infty_c(X,V(\gamma))$ (resp. $C_c^\infty(Y,V_Y(\gamma))$). Note that $\Omega_G$ is elliptic and formally selfadjoint. Using this it is easy to see that the algebra ${\cal Z}_\gamma$ can be generated by essentially selfadjoint elements with commuting resolvents. Thus there exist spectral decompositions of $L^2(X,V(\gamma))$ (resp. $L^2(Y,V_Y(\gamma))$) with respect to ${\cal Z}_\gamma$ (which can equivalently be considered as a spectral decompositions with respect to ${\cal Z}$ as we will do). We will also consider spectral decompositions of these Hilbert spaces with respect to finite abelian extensions of ${\cal Z}_\gamma$ which are obtained by adjoining further essentially selfadjoint differential operators. Since ${\cal Z}_\gamma$ is a quotient of ${\cal Z}$ we can parametrize characters of ${\cal Z}_\gamma$ using the Harish-Chandra isomorphism. Let ${\bf h}\subset {\bf g}$ denote a Cartan algebra of ${\bf g}$, $W=W({\bf g},{\bf h})$ the Weyl group, and ${\bf h}_{\bf C}^*$ the complexified dual of ${\bf h}$. The Harish-Chandra isomorphism identifies characters of ${\cal Z}$ with points in ${\bf h}^*_{\bf C}/W$. Let $\lambda\in{\bf h}^*_{\bf C}$ represent some $W$-orbit. Then we denote the corresponding character of ${\cal Z}$ by $\chi_\lambda$. The abstract spectral decomposition gives a measurable field of Hilbert spaces $\{H_\lambda\}_{\lambda\in {\bf h}^*_{\bf C}/W}$, a measure $\kappa$ on ${\bf h}^*_{\bf C}/W$, and an isometry $$\alpha:L^2(Y,V_Y(\gamma))\cong \int_{{\bf h}^*_{\bf C}/W} H_\lambda \kappa(d\lambda)\ .$$ If we let ${\cal Z}$ act on $H_\lambda$ by the character $\chi_\lambda$, then $\alpha$ is compatible with the action of ${\cal Z}$. It is of course apriori known that $\kappa$ is supported on the set of $\lambda$ with the property that $\chi_\lambda$ factors through ${\cal Z}_\gamma$. We want to describe this set in greater detail. The sphere bundle of $X$ can be identified with the homogeneous space $G/M$, where $M\subset K$. Let ${\bf m}$ denote the Lie algebra of $M$. We choose a Cartan algebra ${\bf t}$ of ${\bf m}$ and let ${\bf a}$ be a one-dimensional subspace of the orthogonal complement of ${\bf k}$ in ${\bf g}$. Then ${\bf a}\oplus{\bf t}=:{\bf h}$ is a Cartan algebra of ${\bf g}$. We further choose a positive root system of ${\bf t}$. Let $\rho_m$ denote half of the sum of the positive roots of $({\bf m},{\bf t})$. For $\sigma\in \hat{M}$ let $\mu_\sigma\in{\bf t}^*$ be its highest weight. Let ${\aaaa_\C^\ast}$ denote the complexification of the dual of ${\bf a}$. A pair $(\sigma\in \hat{M},\lambda\in{\aaaa_\C^\ast})$ determines the character $\chi_{\mu_\sigma+\rho_m-\lambda}$ of ${\cal Z}$. Then representation theory of $G$ implies that $${\mbox{\rm supp}}(\kappa)\subset \{\chi_{\mu_\sigma+\rho_m-\lambda}|[\gamma_{|M}:\sigma]\not=0,\lambda\in{\aaaa_\C^\ast}\}\ .$$ In fact there are more restrictions since characters contributing to the spectral decomposition must be selfadjoint. This restriction implies that if $\chi_{\mu_\sigma+\rho_m-\lambda}\in{\mbox{\rm supp}}(\kappa)$, then $\lambda$ has to be either real or imaginary. Thus we apriori know that the support of $\kappa$ is contained in the projection to ${\bf h}^*_{\bf C}/W$ of a finite union of lines in ${\bf h}^*_{\bf C}$. Speaking about the absolute-continuous part of the spectrum, we have in mind that the measure $\kappa$ restricted to the corresponding one-dimensional set is absolute continuous with respect to the one-dimensional Lebesgue measure. We do not employ this apriory knowledge about the support of $\kappa$ in the proofs, it rather follows from our arguments. The goal of this paper is to describe in detail spectral decomposition of $L^2(Y,V_Y(\gamma))$ with respect to the algebra ${\cal Z}$. The Eisenstein series is used to identify the absolute continuous spectrum. We then show that $\kappa$ has no singular continuous component. Finally obtain the finiteness of the point spectrum and a description of all infinite-dimensional eigenspaces. In analogy to the spectral decomposition in the finite volume case we obtain a decomposition $$L^2(Y,V_Y(\gamma))=L^2(Y,V_Y(\gamma))_c\oplus L^2(Y,V_Y(\gamma))_{res}\oplus L^2(Y,V_Y(\gamma))_{cusp}\oplus L^2(Y,V_Y(\gamma))_{scat} .$$ Here $L^2(Y,V_Y(\gamma))_c$ is the continuous part given by wave packets of Eisenstein series. The space $L^2(Y,V_Y(\gamma))_{res}$ is the finite-dimensional residual part which is essentially generated by the residues of Eisenstein series. The cuspidal part $L^2(Y,V_Y(\gamma))_{cusp}$ consists of a finite number of infinite dimensional eigenspaces and is related to the discrete series representations of $G$ occuring in $L^2(X,V(\gamma))$. The scattering part $L^2(Y,V_Y(\gamma))_{scat}$ consists of finite-dimensional eigenspaces at the boundary of the continuous spectrum. In contrast to the finite-volume case this part can be non-trivial as we demonstrate by an example. The motivation for studying the spectral decomposition with respect to ${\cal Z}$ (and larger commutative algebras) instead of $\Delta$ is that these algebras encode additional symmetries. If only the spectral decomposition of $L^2(Y,V_Y(\gamma))$ with respect to the Laplacian $\Delta$ is considered, then one encounters embedded eigenvalues. Their "stability" is explained by the additional symmetries since they are isolated with respect to the larger algebras. For locally symmetric manifolds of the sort considered in present paper the spectral decomposition of $L^2(Y,V_Y(\gamma))$ with respect to the Laplacian (respectively partial results) were obtained by \begin{itemize} \item Patterson \cite{patterson75} for trivial $\gamma$ and surfaces \item Lax-Phillips \cite{laxphillips82}, \cite{laxphillips841}, \cite{laxphillips842}, \cite{laxphillips85}, Perry \cite{perry87} for higher dimensional hyperbolic manifolds and trivial $\gamma$ \item Mazzeo-Phillips \cite{mazzeophillips90} for differential forms on hyperbolic manifolds \item Epstein-Melrose-Mendoza \cite{epsteinmelrosemendoza91}, Epstein-Melrose \cite{epsteinmelrose90} for differential forms on complex-hyperbolic manifolds. \end{itemize} There is related work on Eisenstein series and the scattering matrix in the real hyperbolic case for trivial $\gamma$ (e.g. \cite{patterson76}, \cite{patterson761}, \cite{patterson89}, \cite{mandouvalos86}, \cite{mandouvalos89}, \cite{perry89}). At the end of this introduction let us make some remarks concerning the methods. Once the Eisenstein series are constructed the realization of the absolute continuous part of the spectrum is almost standard. One of the most important steps in proving the spectral decomposition is to show the absence of the singular continuous spectrum. Usually, the limiting absorption principle (e.g. \cite{perry87}) or commutator methods (see e.g. \cite{froesehislopperry91}) are employed at this point. Here we use a completely different method (proposed in \cite{bernstein88}) which is based on an apriori knowledge of all relevant generalized eigenfunctions of ${\cal Z}$. Our discussion of the point spectrum is based on the asymtotic expansion of eigenfunctions and boundary value theory. Before starting with the main topic of the paper in Section \ref{esess} we analyse the boundary values of generalized eigenfunctions along the geodesic boundary $\partial X$ of $X$. In particular we are interested in the space of $\Gamma$-invariant distributional sections of homogeneous bundles over $\partial X$ with support in the limit set of $\Gamma$. Sections \ref{fiert} to \ref{invvv} are devoted to the analysis on $\partial X$. Our results also have a more representation theoretic interpretation. Let $L^2(\Gamma\backslash G)_K$ denote the space of all $K$-finite vectors on $L^2(\Gamma\backslash G)$. Then combining our results for all $\gamma\in K$ one can obtain a decomposition of $L^2(\Gamma\backslash G)_K$ into unitarizable $({\bf g},K)$-modules and a decomposition of $L^2(\Gamma\backslash G)$ as a direct integral of unitary representations of $G$. The classification of the unitary dual of $G$ then leads to further restrictions of the location of the residual part of the spectrum. \noindent {\it Acknowledgement: We thank R. Mazzeo and P. Perry for discussing of parts of this work.} \section{Geometric preparations}\label{fiert} Let $G$ be a connected, linear, real semisimple Lie group of rank one, $G=KAN$ be an Iwasawa decomposition of $G$, ${\bf g}={\bf k}\oplus{\bf a}\oplus{\bf n}$ be the corresponding Iwasawa decomposition of the Lie algebra ${\bf g}$, $M:=Z_K(A)$ be the centralizer of $A$ in $K$ and $P:=MAN$ be a minimal parabolic subgroup. The group $G$ acts isometrically on the rank-one symmetric space $X:=G/K$. Let $\partial X:=G/P=K/M$ be its geodesic boundary. We consider $X\cup\partial X$ as a compact manifold with boundary. By the classification of symmetric spaces with strictly negative sectional curvature $X$ is one of the following spaces: \begin{itemize} \item a real hyperbolic space, \item a complex hyperbolic space, \item a quaternionic hyperbolic space, \item or the Cayley hyperbolic plane, \end{itemize} and $G$ is a linear group finitely covering of the orientation-preserving isometry group of $X$. Let $\Gamma \subset G$ be a torsion-free, discrete subgroup. \begin{ass}\label{asss} We assume that there is a $\Gamma$-invariant partition $\partial X =\Omega\cup \Lambda$, where $\Omega\not=\emptyset$ is open and $\Gamma$ acts freely and cocompactly on $X\cup\Omega$. \end{ass} The locally symmetric space $Y:=\Gamma\backslash X$ is a complete Riemannian manifold of infinite volume without cusps. It can be compactified by adjoining the geodesic boundary $B:=\Gamma\backslash \Omega$. We call $\Lambda$ the limit set of $\Gamma$. A group $\Gamma$ satisfying \ref{asss} is also called convex-cocompact since it acts cocompactly on the convex hull of the limit set $\Lambda$. The quotient $Y$ can be called a Kleinian manifold in generalizing the corresponding notion for three-dimensional hyperbolic manifolds. We now consider some geometric consequences of \ref{asss} which eventually allow us to define the exponent $\delta_\Gamma$ of $\Gamma$. Let $g=\kappa(g)a(g)n(g)$, $\kappa(g)\in K$, $a(g)\in A$, $n(g)\in N$ be defined with respect to the given Iwasawa decomposition. By ${\aaaa_\C^\ast}$ we denote the comlexified dual of ${\bf a}$. If $\lambda\in {\aaaa_\C^\ast}$, then we set $a^\lambda:={\rm e}^{\langle \lambda,\log(a)\rangle}\in{\bf C}$. The roots of ${\bf a}$ corresponding to ${\bf n}$ distinguish a positive cone ${\bf a}^*_+$. Define $\rho\in {\bf a}_+^*$ as usual by $\rho(H):=\frac{1}{2}{\mbox{\rm tr}}({\mbox{\rm ad}}(H)_{|{\bf n}})$, $\forall H\in{\bf a}$. We adopt the following conventions about the notation for points of $X$ and $\partial X$. A point $x\in \partial X$ can equivalently be denoted by a subset $kM\subset K$ or $gP\subset G$ representing this point in $\partial X=K/M$ or $\partial X=G/P$. If $F\subset \partial X$, then $FM:=\bigcup_{kM\in F}kM\subset K$. Analogously, we can denote $b\in X$ by $gK\subset G$, where $gK$ represents $b$ in $X=G/K$. \begin{lem}\label{no} For any compact $F\subset \Omega$ we have $\sharp(\Gamma\cap FMA_+K)<\infty$. \end{lem} {\it Proof.$\:\:\:\:$} The compact set $FMA_+K\cup F\subset X\cup\Omega$ contains at most a finite number of points of the orbit $\Gamma K$ of the origin of $X$. \hspace*{\fill}$\Box$ \\[0.5cm]\noindent Let $A_+:=\exp({\bf a}_+)$. Any element $g\in G$ has a decomposition $g=ka_gh$, $k,h\in K$, $a_g\in A_+\cup \{1\}$, where $a_g$ is uniquely determined by $g$. \begin{lem}\label{no1} Let $k_0M \in \partial X$. For any compact $W\subset (\partial X\setminus k_0M)M$ there exists a neighbourhood $U\subset K$ of $k_0M$ and constants $c>0$, $C<\infty$, such that for all $g=ha_gh^\prime \in WA_+K$ and $k\in U$ \begin{equation}\label{unm} c a_g \le a(g^{-1}k) \le C a_g \ . \end{equation} \end{lem} {\it Proof.$\:\:\:\:$} The set $W^{-1}k_0M$ is compact and disjoint from $M$. Let $w\in N_K(M)$ represent the non-trivial element of the Weyl group of $({\bf g},{\bf a})$. Set $\bar{{\bf n}}=\theta({\bf n})$, where $\theta$ is the Cartan involution of $G$ fixing $K$ and define $\bar{N}:=exp(\bar{{\bf n}})$. There is a precompact open $V\subset \bar{N}$ such that $W^{-1}k_0M\subset w \kappa(V)M$. By enlarging $V$ we can assume that $V$ is $A_+$-invariant, where $A$ acts on $\bar{N}$ by $(a,\bar{n})\mapsto a\bar{n}a^{-1}$. Moreover, there exists an open neighbourhood $U\subset K$ of $k_0M$ such that $w^{-1}W^{-1}U M\subset \kappa(V)M$. Let $k\in U$ and $g=ha_gh^\prime \in WA_+K$. Then we have $h^{-1}k=w\kappa(\bar{n})m $ for $\bar{n}\in V$, $m\in M$. Furthermore, \begin{eqnarray*} a(g^{-1}k)&=&a(h^{\prime -1}a_g^{-1}h^{-1}k)\\ &=& a(a_g^{-1}w\kappa(\bar{n})m) \\ &=&a(a_g \kappa(\bar{n}))\\ &=&a(a_g\bar{n}n(\bar{n})^{-1}a(\bar{n})^{-1})\\ &=&a(a_g\bar{n}a_g^{-1}) a(\bar{n})^{-1} a_g\ . \end{eqnarray*} Now $a_g\bar{n}a_g^{-1}\in V$. Set \begin{eqnarray*} c&:=&\inf_{\bar{n}\in V} a(\bar{n}) \inf_{\bar{n}\in V} a(\bar{n})^{-1}\\ C&:=&\sup_{\bar{n}\in V} a(\bar{n}) \sup_{\bar{n}\in V} a(\bar{n})^{-1}\ . \end{eqnarray*} Since $V$ is precompact we have $0 < c \le C<\infty$. It follows that $c a_g \le a(g^{-1}k )\le Ca_g\ .$ \hspace*{\fill}$\Box$ \\[0.5cm]\noindent \begin{lem}\label{poi} The series $$\sum_{g\in\Gamma} a_g^{-2\rho}$$ converges. \end{lem} {\it Proof.$\:\:\:\:$} We represent $\partial X=G/P$. There is a $P$-invariant splitting ${\bf g}=\bar{{\bf n}}\oplus{\bf p}$ and an identification $\bar{{\bf n}}^*={\bf n}$ by the invariant bilinear form on ${\bf g}$. In particular $\Lambda^{max}\bar{{\bf n}}^*=\Lambda^{max}{\bf n}$ as a $P$-module. It follows that \begin{equation}\label{sad}\Lambda^{max}T^*\partial X=G\times_P \Lambda^{max}{\bf n} \ . \end{equation} This bundle is trivial as a $K$-homogeneous bundle. Fix an orientation of $\partial X$ and let $\omega\in C^\infty(\partial X, \Lambda^{max}T^*\partial X)$ be a positive $K$-invariant volume form, which is unique up to a positive scalar factor. Let $g\in G$. Then $g$ acts as a diffeomorphism on $\partial X$. Using (\ref{sad}) and the $K$-invariance of $\omega$ we obtain $(g^*\omega)(kM)=a(g^{-1}k)^{-2\rho}\omega(kM)$. Let $F\subset \Omega$ be a compact set with non-trivial interior such that $gF\cap F=\emptyset$ for all $1\not=g\in\Gamma$. Such $F$ exists by Assumption \ref{asss}. Then \begin{eqnarray*} \infty&>& \int_{\partial X} \omega > \int_{\cup_{g\in\Gamma}\: gF} \omega\\ &=& \sum_{g\in\Gamma} \int_{gF} \omega = \sum_{g\in\Gamma} \int_F g^*\omega\\ &=& \sum_{g\in\Gamma} \int_F a(g^{-1}k)^{-2\rho} \omega(kM). \end{eqnarray*} Let $F_1\subset \Omega$ be a compact neighbourhood of $F$. By Lemma \ref{no} the set $\Gamma\cap F_1MA_+K$ is finite. We apply the Lemma \ref{no1} taking for $W$ the closure of $\Gamma\setminus (\Gamma\cap F_1MA_+K)$. Then we can cover $FM$ with finitly many sets $U$ the existence of which was asserted in that lemma. Thus here is a constant $C\in A$ such that for all $g\in \Gamma\setminus (\Gamma\cap F_1MA_+K)$ and $k\in FM$ $$a(g^{-1}k)\le Ca_g\ .$$ It follows that \begin{eqnarray*} \sum_{g\in\Gamma} a_g^{-2\rho} \int_F \omega&=& \sum_{g\in\Gamma} \int_F a_g^{-2\rho} \omega\\& \le& C^{2\rho} \sum_{g\in\Gamma\setminus(\Gamma\cap F_1 MA_+K)} \int_F a(g^{-1}kM)^{-2\rho} \omega(kM)\ \ + \sum_{g\in\Gamma\cap F_1MA_+K} a_g^{-2\rho}\\[0.5cm] &<& \infty\ . \end{eqnarray*} This implies the lemma since $$ \int_F \omega\not=0\ .$$ \hspace*{\fill}$\Box$ \\[0.5cm]\noindent \begin{ddd} Let $\delta_\Gamma\in {\bf a}^*$ be the smallest element such that $\sum_{g\in\Gamma} a_g^{- \rho-\lambda}$ converges for all $\lambda\in{\bf a}^*$ with $\lambda>\delta_\Gamma$. \end{ddd} $\delta_\Gamma$ is called the exponent of $\Gamma$. By Proposition \ref{poi} we have $\delta_\Gamma \le \rho$. If $\Gamma$ is non-trivial, then $\delta_\Gamma\ge -\rho$. For the trivial group we have $\delta_{\{1\}}=-\infty$. If $X$ is the $n$-dimensional hyperbolic space and we identify ${\bf a}^*$ with ${\bf R}$ such that $\rho=\frac{n-1}{2}$, then it was shown by Patterson \cite{patterson762} and Sullivan \cite{sullivan79}, that $\delta_\Gamma+\frac{n-1}{2}=\dim_H(\Lambda)$, where $\dim_H$ denotes the Hausdorff dimension (the Hausdorff dimension of the empty set is by definition $-\infty$). It was also shown in \cite{sullivan79} that in the real hyperbolic case $\sum_{g\in\Gamma} a_g^{- \rho-\delta_\Gamma}$ diverges. Hence $\delta_\Gamma<\rho$. In the proof of the meromorphic continuation of the Eisenstein series we employ at a certain place that $X$ belongs to a series of symmetric spaces with increasing dimensions. It is there where we need the following assumption \begin{ass}\label{caly} If $X$ is the Cayley hyperbolic plane, then we assume that $\delta_\Gamma<0$. \end{ass} We believe that this assumption is only of technical nature and can be dropped using other methods for the meromorphic continuation of the Eisenstein series. \section{Analytic preparations}\label{anaprep} The goal of the following two sections is to construct $\Gamma$-invariant vectors in principal series representations of $G$. For this reason we introduce the extension map $ext$ and the scattering matrix, and construct their meromorphic continuations. The principal series representations of $G$ are realized on spaces of distribution sections of bundles over $\partial X$. Roughly speaking the extension map extends a $\Gamma$-invariant distribution sections on $\Omega$ across the limit set $\Lambda$. The space of $\Gamma$-invariant distributions on $\Omega$ is easily described as the space of distribution sections on bundles over $B$. The majority of the $\Gamma$-invariant vectors of the principal series is then obtained by extension. Consider a finite-dimensional unitary representation $\sigma$ of $M$ on $V_\sigma$. If $w\in W({\bf g},{\bf a})$ is the non-trivial element of the Weyl group, then we can form the representation $\sigma^w$ of $M$ on $V_\sigma$ by conjugating the argument of $\sigma$ with a representative of $w$ in $N_K(M)$. In this section we will assume that $\sigma$ is irreducible. Note that we will change this convention later in the case that $\sigma$ and $\sigma^w$ are non-equivalent by considering the sum $\sigma\oplus \sigma^w$ instead. For $\lambda\in {\aaaa_\C^\ast}$ we form the representation $\sigma_\lambda$ of $P$ on $V_{\sigma_\lambda}:=V_\sigma$, which is given by $\sigma_\lambda(man):=\sigma(m)a^{\rho-\lambda}$. Let $V(\sigma_\lambda):=G\times_P V_{\sigma_\lambda}$ be the associated homogeneous bundle. Set $V_B(\sigma_\lambda):=\Gamma\backslash V(\sigma_\lambda)$. Let $\tilde{\sigma}$ be the dual representation to $\sigma$. Then there are natural pairings \begin{eqnarray*} V(\tilde{\sigma}_{-\lambda})\otimes V(\sigma_\lambda)&\rightarrow& \Lambda^{max}T^*\partial X\\ V_B(\tilde{\sigma}_{-\lambda})\otimes V_B(\sigma_\lambda)&\rightarrow& \Lambda^{max}T^*\partial B\ . \end{eqnarray*} The orientation of $\partial X$ induces one of $B$. Employing these pairings and integration with respect to the fixed orientation we obtain identifications \begin{eqnarray*} C^{-\infty}(\partial X,V(\sigma_\lambda))&=&C^{\infty}(\partial X,V(\tilde{\sigma}_{-\lambda}))^\prime\\ C^{-\infty}(B,V_B(\sigma_\lambda))&=&C^{\infty}(B,V_B(\tilde{\sigma}_{-\lambda}))^\prime\ . \end{eqnarray*} As a $K$-homogeneous bundle we have a canonical identification $V(\sigma_\lambda)\cong K\times_M V_\sigma$. Thus $\bigcup_{\lambda\in{\aaaa_\C^\ast}} V(\sigma_\lambda)\rightarrow {\aaaa_\C^\ast}\times \partial X$ has the structure of a trivial holomorphic family of bundles. Let $\pi^{\sigma,\lambda}$ denote the representation of $G$ on the space of sections of $V(\sigma_\lambda)$ given by the left-regular representation. Then $\pi^{\sigma,\lambda}$ is called a principal series representation of $G$. Note that there are different globalizations of this representation which are distinguished by the regularity of the sections (smooth, distribution e.t.c.). For any small open subset $U\subset B$ and diffeomorphic lift $\tilde{U}\subset \Omega$ the restriction $V_B(\sigma_\lambda)_{|U}$ is canonically isomorphic to $V(\sigma_\lambda)_{|\tilde{U}}$. Let $\{U_\alpha\}$ be a cover of $B$ by open sets as above. Then $$\bigcup_{\lambda\in{\aaaa_\C^\ast}} V_B(\sigma_\lambda)\rightarrow {\aaaa_\C^\ast}\times B$$ can be given the structure of a holomorphic family of bundles by glueing the trivial families $$\bigcup_{\lambda\in{\aaaa_\C^\ast}}V_B(\sigma_\lambda)_{|U}\cong \bigcup_{\lambda\in{\aaaa_\C^\ast}}V(\sigma_\lambda)_{|\tilde{U}}$$ together using the holomorphic families of glueing maps induced by $\pi^{\sigma,\lambda}(g)$, $g\in\Gamma$. Thus it makes sense to speak of holomorphic or smooth or continuous families of sections ${\aaaa_\C^\ast}\ni\mu\mapsto f_\mu\in C^{\pm\infty}(B,V_B(\sigma_\mu))$. When dealing with holomorphic families of vectors in topological vector spaces we will employ the following functional analytic facts. Let ${\cal F},{\cal G},{\cal H} \dots$ be complete locally convex topological vector spaces. A locally convex vector space is called a Montel space if its closed bounded subsets are compact. A Montel space is reflexive, i.e., the canonical map into its bidual is an isomorphism. Moreover, the dual space of a Montel space is again a Montel space. \begin{fact} The space of smooth sections of a vector bundle and its topological dual are Montel spaces. \end{fact} We equip ${\mbox{\rm Hom}}({\cal F},{\cal G})$ with the topology of uniform convergence on bounded sets. Let $V\subset {\bf C}$ be open. A map $f:V\rightarrow {\mbox{\rm Hom}}({\cal F},{\cal G})$ is called holomorphic if for any $z_0\in V$ there is a sequence $f_i\in {\mbox{\rm Hom}}({\cal F},{\cal G})$ such that $f(z)=\sum_{n=0}^\infty f_i (z-z_0)^i$ converges for all $z$ close to $z_0$. Let $f:V\setminus \{z_0\} \rightarrow {\mbox{\rm Hom}}({\cal F},{\cal G})$ be holomorphic and $f(z)=\sum_{n=-N}^\infty f_i (z-z_0)^i$ for all $z\not=z_0$ close to $z_0$. Then we say that $f$ is meromorphic and has a pole of order $N$ at $z_0$. If $f_i$, $i=-N,\dots,-1$, are finite dimensional, then $f$ has, by definition, a finite-dimensional singularity. We call a subset $A\subset {\cal F}\times {\cal G}^\prime$ sufficient if for $B\in {\mbox{\rm Hom}}({\cal F},{\cal G})$ the condition $<\phi,B \psi >=0$, $ \forall (\psi,\phi)\in A$, implies $B=0$. \begin{fact}\label{holla} The following assertions are equivalent : \begin{enumerate} \item (i) $\:\:f:V\rightarrow {\mbox{\rm Hom}}({\cal F},{\cal G})$ is holomorphic. \item (ii) $\:\:\: f$ is continuous and there is a sufficient set $A\subset {\cal F}\times {\cal G}^\prime$ such that for all $(\psi,\phi)\in A$ the function $V\ni z\mapsto \langle \phi,f(z)\psi\rangle $ is holomorphic. \end{enumerate} \end{fact} \begin{fact}\label{seq} Let $f_i:V\rightarrow Hom({\cal F},{\cal G})$ be a sequence of holomorphic maps. Moreover let $f :V\rightarrow Hom({\cal F},{\cal G})$ be continuous such that for a sufficient set $A \subset {\cal F}\times {\cal G}^\prime$ the functions $\langle \phi,f_i \psi\rangle $, $(\psi,\phi)\in A$, converge locally uniformly in $V$ to $\langle\phi,f \psi\rangle$. Then $f$ is holomorphic, too. \end{fact} \begin{fact}\label{adjk} Let $f:V\rightarrow {\mbox{\rm Hom}}({\cal F},{\cal G})$ be continuous. Then the adjoint $f^\prime:V\rightarrow {\mbox{\rm Hom}}({\cal G}^\prime,{\cal F}^\prime)$ is continuous. If $f$ is holomorphic, then so is $f^\prime$. \end{fact} \begin{fact}\label{comp} Assume that ${\cal F}$ is a Montel space. Let $f:V\rightarrow {\mbox{\rm Hom}}({\cal F},{\cal G})$ and $f_1:V\rightarrow {\mbox{\rm Hom}}({\cal G},{\cal H})$ be continuous. Then $f_1\circ f : V\rightarrow {\mbox{\rm Hom}}({\cal F},{\cal H})$ is continuous. If $f,f_1$ are holomorphic, so is $f_1\circ f$. \end{fact} The following lemma will be employed in Section \ref{extttttt}. Since it is of purely functional analytic nature we consider it at this place. Let ${\cal H}$ be a Hilbert space and ${\cal F}\subset {\cal H}$ be a Fr\'echet space such that the embedding is continuous and compact. In the application we have in mind ${\cal H}$ will be some $L^2$- space of sections of a vector bundle over a compact closed manifold and ${\cal F}$ be the Fr\'echet space of smooth sections of this bundle. The continuous maps ${\mbox{\rm Hom}}({\cal H},{\cal F})$ will be called smoothing operators. Let $V\subset {\bf C}$ be open and connected, and $V\ni z\rightarrow R(z)\in {\mbox{\rm Hom}}({\cal H},{\cal F})$ be a meromorphic family of smoothing operators with at most finite-dimensional singularities. Note that $R(z)$ is a meromorphic family of compact operators on ${\cal H}$ in a natural way. \begin{lem} \label{merofred} If $1-R(z)$ is invertible at some point $z\in V$ where $R(z)$ is regular, then $$(1-R(z))^{-1}=1-S(z)\ ,$$ where $V\ni z\rightarrow S(z)\in {\mbox{\rm Hom}}({\cal H},{\cal F})$ is a meromorphic family of smoothing operators with at most finite dimensional singularities. \end{lem} {\it Proof.$\:\:\:\:$} We apply Reed-Simon IV \cite{reedsimon78}, Theorem XIII.13 in order to conclude that $(1-R(z))^{-1}$ is a meromorphic family of operators on ${\cal H}$ having at most finite-dimensional singularities. Making the ansatz $(1-R(z))^{-1}= 1-S(z)$, where apriori $S(z)$ is a meromorphic familiy of bounded operators on ${\cal H}$ with finite dimensional singularities, we obtain $S=-R-R\circ S$. This shows that $S$ is a meromorpic family in ${\mbox{\rm Hom}}({\cal H},{\cal F})$. \hspace*{\fill}$\Box$ \\[0.5cm]\noindent This finishes the functional analytic preparations and we now turn to the construction of the extension map. In fact, we first introduce its adjoint which is the push-down $$\pi_\ast:C^\infty(\partial X,V(\sigma_\lambda))\rightarrow C^\infty(B,V_B(\sigma_\lambda))\ .$$ Using the identification $C^\infty(B,V_B(\sigma_\lambda))= {}^\Gamma C^\infty(\Omega,V(\sigma_\lambda))$ we define $\pi_\ast$ by \begin{equation}\label{summ}\pi_*(f)(kM)= \sum_{g\in\Gamma} (\pi(g)f)(kM),\quad kM\in\Omega \ ,\end{equation} if the sum converges. Here $\pi(g)$ is the action induced by $\pi^{\sigma,\lambda}(g)$. Note that the universal enveloping algebra ${\cal U}({\bf g})$ is a filtered algebra. Let ${\cal U}({\bf g})_m$, $m\in{\bf N}_0$, be the space of elements of degree less or equal than $m$. For any $m$ and bounded subset $A\subset {\cal U}({\bf g})_m$ we define the seminorm $\rho_{m,A}$ on $C^\infty(\partial X,V(\sigma_\lambda))$ by $$\rho_{m,A}(f):=\sup_{X\in A, k\in K}|f(\kappa(kX))|\ .$$ These seminorms define the Fr\'echet topology of $C^\infty(\partial X,V(\sigma_\lambda))$ (in fact a countable set of such seminorms is sufficient). In order to define the Fr\'echet topology on $C^\infty(B,V_B(\sigma_\lambda))$ we fix an open cover $\{U_\alpha\}$ of $B$ such that each $U_\alpha$ has a diffeomorphic lift $\tilde{U}_\alpha \subset\Omega$. Then we have canonical isomorphisms $$C^\infty(\tilde{U}_\alpha ,V(\sigma_\lambda))\cong C^\infty(\ U_\alpha,V_B(\sigma_\lambda))\ .$$ For any $U\in \{U_\alpha\}$ we define the topology of $C^\infty(\tilde{U},V(\sigma_\lambda))$ using the seminorms $$\rho_{U,m,A}(f):=\sup_{X\in A, k\in \tilde{U}M }|f(\kappa(kX))|\ ,$$ where $m\in{\bf N}_0$ and $A\subset {\cal U}({\bf g})_m$ is bounded. Since $C^\infty(B,V_B(\sigma_\lambda))$ maps to $C^\infty( U_\alpha,V_B(\sigma_\lambda))$ by restriction for each $\alpha$ we obtain a system of seminorms defining the Fr\'echet topology of $C^\infty(B,V_B(\sigma_\lambda))$. \begin{lem}\label{anal1} If ${\rm Re }(\lambda)<-\delta_\Gamma$, then the sum (\ref{summ}) converges for $f\in C^\infty(\partial X,V(\sigma_\lambda))$ and defines a continuous map $$\pi_\ast: C^\infty(\partial X,V(\sigma_\lambda))\rightarrow C^\infty(B,V_B(\sigma_\lambda))\ .$$ Moreover, $\pi_*$ depends holomorphically on $\lambda$. \end{lem} {\it Proof.$\:\:\:\:$} Consider $U\in \{U_\alpha\}$. We want to estimate $$C^\infty(\partial X,V(\sigma_\lambda))\ni f\mapsto res_{\tilde{U}}\circ \pi(g) f \in C^\infty(\tilde{U},V(\sigma_\lambda))\ .$$ Let $\Delta:{\cal U}({\bf g})\rightarrow {\cal U}({\bf g})\otimes {\cal U}({\bf g})$ be the coproduct and write $\Delta(X)=\sum_i X_i\otimes Y_i$. Fix $l\in {\bf N}_0$ and a bounded set $A\in {\cal U}({\bf g})_l$. Then there is another bounded set $A_1\subset {\cal U}({\bf g})_l$ depending on $A$ such that \begin{eqnarray*} \rho_{U,l,A}(res_{\tilde{U}M}\circ \pi(g) f)&=&\sup_{X\in A,k\in \tilde{U}}|(\pi(g)f)(\kappa(kX))|\\ &=& \sup_{X\in A,k\in \tilde{U}M}|\sum_ i a(g^{-1}\kappa(kX_i))^{\lambda-\rho}f(\kappa(g^{-1}\kappa(kY_i)))|\\ &\le & \sup_{X\in A_1,k\in \tilde{U}M}| a(g^{-1}kX)^{\lambda-\rho}| \sup_{X\in A_1,k\in \tilde{U}M}| f(\kappa(g^{-1}kX))| \ . \end{eqnarray*} The Poincar\'e-Birkhoff-Witt theorem gives a decomposition ${\cal U}({\bf g})={\cal U}(\bar{{\bf n}}){\cal U}({\bf m}){\cal U}({\bf a})\oplus {\cal U}({\bf g}){\bf n}$. Let $q:{\cal U}({\bf g})\rightarrow {\cal U}(\bar{{\bf n}}){\cal U}({\bf m}){\cal U}({\bf a})$ be the associated projection. Then for $g\in G$ and $X\in{\cal U}({\bf g})$ we have $\kappa(gX)=\kappa(gq(X))$, $a(gX)=a(gq(X))$. Let $U_1\subset \Omega$ be an open neighbourhood of $\tilde{U}$. Then by Lemma \ref{no} the intersection $\Gamma\cap U_1MA_+K$ is finite. Let $W:=(\partial X\setminus U_1)M$. Then by Lemma \ref{no1} we can find a compact $A_+$-invariant set $V\subset \bar{{\bf n}}$ such that $W^{-1}\tilde{U}M\subset w\kappa(V)M$. For $g=ha_gh^\prime\in WA_+K$ and $k\in \tilde{U}M$ we obtain $h^{-1}k=w\kappa(\bar{n})m$ for some $\bar{n}\in V$, $m\in M$. Let $X\in {\cal U}({\bf g})$. Then \begin{eqnarray*} \kappa(g^{-1}kX)&=&\kappa(h^{\prime -1}a_g^{-1}h^{-1}kX)\\ &=&h^{\prime -1}\kappa(a_g^{-1}w\kappa(\bar{n})mX)\\ &=&h^{\prime -1}w\kappa(a_g\bar{n}n(\bar{n})^{-1}a(\bar{n})^{-1}mX) \\ &=&h^{\prime -1}w\kappa(a_g\bar{n}a_g^{-1} a_g[n(\bar{n})^{-1}a(\bar{n})^{-1}mXm^{-1}a(\bar{n})n(\bar{n})]a_g^{-1})m\\ &=&h^{\prime -1}w\kappa(a_g\bar{n}a_g^{-1} a_gq(n(\bar{n})^{-1}a(\bar{n})^{-1}mXm^{-1}a(\bar{n})n(\bar{n}))a_g^{-1})m\ . \end{eqnarray*} Since $V$ is compact the sets $n(V)^{-1}a(V)^{-1}MA_1Ma(V)n(V)=:A_2\subset {\cal U}({\bf g})_l$ and $q(A_2)$ are bounded. Conjugating $q(A_2)$ with $A_+$ gives clearly another bounded set $A_3\subset {\cal U}({\bf g})_l$. We can find a bounded set $A_4\subset {\cal U}({\bf g})_l$ such that $\kappa(a_g\bar{n}a_g^{-1}A_3)\subset \kappa(\kappa(a_g\bar{n}a_g^{-1})A_4)$ for all $a_g\in A_+$. This implies for $g\in WA_+K$ that \begin{equation}\label{kkll1}\sup_{X\in A_1,k\in \tilde{U}}| f(\kappa(g^{-1}kX))|\le \rho_{l,A_4}(f)\ .\end{equation} We also have \begin{eqnarray*} a(g^{-1}kX)&=&a(h^{\prime -1}a_g^{-1}h^{-1}kX)\\ &=& a(a_g^{-1}w\kappa(\bar{n})mX)\\ &=&a(a_g \kappa(\bar{n})mXm^{-1})\\ &=&a(a_g\bar{n}n(\bar{n})^{-1}a(\bar{n})^{-1}mXm^{-1})\\ &=&a(a_g\bar{n}a_g^{-1}a_g n(\bar{n})^{-1}a(\bar{n})^{-1}mXm^{-1}a(\bar{n})n(\bar{n})a_g^{-1}) a(\bar{n})^{-1} a_g\ .\\ &=&a(a_g\bar{n}a_g^{-1}a_g q(n(\bar{n})^{-1}a(\bar{n})^{-1}mXm^{-1}a(\bar{n})n(\bar{n}))a_g^{-1}) a(\bar{n})^{-1} a_g\ . \end{eqnarray*} Again there is a constant $C<\infty$ such that $$ |a(a_g\bar{n}a_g^{-1}a_g q(n(\bar{n})^{-1}a(\bar{n})^{-1}mXm^{-1}a(\bar{n})n(\bar{n}))a_g^{-1})^{\lambda-\rho}| a(\bar{n})^{ \rho- \lambda } <C$$ for all $a_g\in A_+$, $\bar{n}\in V$, $m\in M$, and $X\in A_1$. It follows that \begin{equation}\label{hj1}\sup_{X\in A_1,k\in \tilde{U}M}| a(g^{-1}kX)^{\lambda-\rho}|\le C a_g^{\lambda-\rho}\end{equation} for almost all $g\in \Gamma$. The estimates (\ref{kkll1}) and (\ref{hj1}) together imply that the sum $$C^l(\partial X,V(\sigma_\lambda))\ni f\mapsto \sum_{g\in\Gamma} res_{\tilde{U}}\circ \pi(g) f \in C^l(\tilde{U},V(\sigma_\lambda))$$ converges for ${\rm Re }(\lambda)<-\delta_\Gamma$ and defines a continuous map of Banach spaces. This map depends holomorphically on $\lambda$ by Fact \ref{seq}. Combining these considerations for all $U\in \{U_\alpha\}$ and $l\in{\bf N}_0$ we obtain that $$\pi_*:C^\infty(\partial X,V(\sigma_\lambda))\rightarrow C^\infty(B,V_B(\sigma_\lambda))$$ is defined and continuous for ${\rm Re }(\lambda)<-\delta_\Gamma$. Moreover it is easy to see that $\pi_*$ depends holomorphically on $\lambda$. \hspace*{\fill}$\Box$ \\[0.5cm]\noindent Still postponing the introduction of the extension map we now consider its left-inverse, the restriction $$res:{}^\Gamma C^{-\infty}(\partial X,V(\sigma_\lambda))\rightarrow C^{-\infty}(B,V_B(\sigma_\lambda))\ .$$ In fact the space ${}^\Gamma C^{-\infty}(\Omega,V(\sigma_\lambda))$ of $\Gamma$-invariant vectors in $C^{-\infty}(\Omega,V(\sigma_\lambda))$ can be canonically identified with the corresponding space $C^{-\infty}(B,V_B(\sigma_\lambda))$. Composing this identification with the restriction $res_\Omega:C^{-\infty}(\partial X,V(\sigma_\lambda))\rightarrow C^{-\infty}(\Omega,V(\sigma_\lambda))$ we obtain the required restriction map $res$. \begin{lem}\label{lopi} There exists a continous map $$\widetilde{res}: C^{-\infty}(\partial X ,V(\sigma_\lambda))\rightarrow C^{-\infty}(B,V_B(\sigma_\lambda))\ ,$$ which depends holomorphically on $\lambda$ and coincides with $res$ on ${}^\Gamma C^{-\infty}(\Omega,V(\sigma_\lambda))$. \end{lem} {\it Proof.$\:\:\:\:$} We exhibit $\widetilde{res}$ as the adjoint of a continuous map $$\pi^*:C^\infty(B,V_B(\tilde{\sigma}_{-\lambda}))\rightarrow C^\infty (\partial X ,V(\tilde{\sigma}_{-\lambda}))$$ which depends holomorphically on $\lambda$. Then the lemma follows from Fact \ref{adjk}. Let $\{U_\alpha\}$ be a finite open cover of $B$ such that each $U_\alpha$ has a diffeomorphic lift $\tilde{U}_\alpha \subset\Omega$. Choose a subordinated partition of unity $\{\chi_\alpha\}$. Pulling $\chi_\alpha$ back to $\tilde{U}_\alpha$ and extending the resulting function by $0$ we obtain a function $\tilde \chi_\alpha\in C^\infty (\partial X ,V(\tilde{\sigma}_{-\lambda}))$. We define $$\pi^*(f):=\sum_\alpha \tilde\chi_\alpha f, \quad f\in C^\infty(B,V_B(\tilde{\sigma}_{-\lambda}))\ ,$$ where we consider $f$ as an element of ${}^\Gamma C^{-\infty}(\partial X,V(\tilde{\sigma}_{-\lambda}))$. Then we set $\widetilde{res}:=(\pi^*)^\prime$. \hspace*{\fill}$\Box$ \\[0.5cm]\noindent The extension map $ext$ will be defined as an right inverse to $res$. \begin{ddd}\label{defofext} For ${\rm Re }(\lambda)>\delta_\Gamma$ we define the extension map $$ext: C^{-\infty}(B,V_B(\sigma_\lambda))\rightarrow {}^\Gamma C^{-\infty}(\partial X,V(\sigma_\lambda))$$ to be the adjoint of $$\pi_*:C^{ *}(\partial X,V(\tilde{\sigma}_{-\lambda})) \rightarrow C^{*}(B,V_B(\tilde{\sigma}_{-\lambda}))\ .$$ \end{ddd} This definition needs a justification. In fact, by Lemma \ref{anal1} the extension exists, is continuous, and by Fact \ref{adjk} it depends holomorphically on $\lambda$. Moreover, it is easy to see that the range of the adjoint of $\pi_*$ consists of $\Gamma$-invariant vectors. \begin{lem}\label{iden} We have $res\circ ext = {\mbox{\rm id}}$. \end{lem} {\it Proof.$\:\:\:\:$} Recall the definition of $\pi^*$ from the proof of Lemma \ref{lopi}. Then $\pi_*\pi^*$ is the identity on $C^\infty(B,V_B(\tilde{\sigma}_{-\lambda}))$. We obtain $$res \circ ext=\widetilde{res}\circ ext= (\pi^*)^\prime \circ (\pi_*)^\prime =(\pi_*\pi^*)^\prime={\mbox{\rm id}}\ .$$ \hspace*{\fill}$\Box$ \\[0.5cm]\noindent Let $C^{-\infty}(\Lambda,V(\sigma_\lambda))$ denote the space of distribution sections of $V(\sigma_\lambda)$ with support in the limit set $\Lambda$. Since $\Lambda$ is $\Gamma$-invariant $C^{-\infty}(\Lambda,V(\sigma_\lambda))$ is a subrepresentation of the representation $\pi^{\sigma,\lambda}$ of $\Gamma$ on $C^{-\infty}(\partial X,V(\sigma_\lambda))$. Note that ${}^\Gamma C^{-\infty}(\Lambda,V(\sigma_\lambda))$ is exactly the kernel of $res$. \begin{lem}\label{mainkor} If ${}^\Gamma C^{-\infty}(\Lambda ,V(\sigma_\lambda))=0$ and if $ext$ is defined, then we have $ext\circ res ={\mbox{\rm id}}$. \end{lem} {\it Proof.$\:\:\:\:$} The assumption implies that $res$ is injective. By Lemma \ref{iden} we have $res(ext\circ res - {\mbox{\rm id}})=0$. \hspace*{\fill}$\Box$ \\[0.5cm]\noindent In order to apply this lemma we have to check its assumption. In the course of the paper we will prove several vanishing results for ${}^\Gamma C^{-\infty}( \Lambda ,V(\sigma_\lambda))$. One of these is available at this point. \begin{lem}\label{pokm} If ${\rm Re }(\lambda)>0$ and ${\rm Im}(\lambda)\not=0$, then ${}^\Gamma C^{-\omega}( \Lambda ,V(\sigma_\lambda))=0$. \end{lem} {\it Proof.$\:\:\:\:$} We employ the Poisson transform. Let $\gamma$ be a finite-dimensional representation of $K$ on $V_\gamma$ such that there exists an injective $T\in {\mbox{\rm Hom}}_M(V_\sigma,V_\gamma)$. We will view sections of $V(\gamma)$ as functions from $G$ to $V_\gamma$ satisfying the usual $K$-invariance condition. \begin{ddd}\label{defofpoi} The Poisson transform $$P:=P^T_\lambda:C^{-\infty}(\partial X,V(\sigma_\lambda))\rightarrow C^\infty(X,V(\gamma))$$ is defined by $$(P \phi)(g):=\int_K a(g^{-1}k)^{-(\rho+\lambda)}\gamma(\kappa(g^{-1}k)) T \phi(k) dk\ .$$ Here $\phi\in C^{-\infty}(\partial X, V(\sigma_\lambda))$ and the integral is a formal notation meaning that the distribution $\phi$ has to be applied to the smooth integral kernel. \end{ddd} In \cite{olbrichdiss} it is shown that the Poisson transform $P$ intertwines the left-regular representations of $G$ on $C^{-\infty}(\partial X,V(\sigma_\lambda))$ and $C^\infty(X,V(\gamma))$. For ${\rm Im}(\lambda)\not=0$ the principal series representation $\pi^{\sigma,\lambda}$ of $G$ on $C^{-\infty}(\partial X, V(\sigma_\lambda))$ is topologically irreducible. Since the Poisson transform $P$ does not vanish and is continuous it is injective. There is a real constant $c_\sigma$ (see \cite{bunkeolbrich955}) such that $(-\Omega_G+c_\sigma+\lambda^2)P \phi =0$ (where $\lambda^2 :=\langle \lambda,\lambda\rangle $ with respect to the ${\bf C}$-linear scalar product induced on ${\aaaa_\C^\ast}$ by the the invariant form on ${\bf g}$). Let $V\subset \partial X$ and $U\subset X$ such that ${\rm clo}(U)\cap V=\emptyset$, where we take the closure of $U$ in $X\cup\Omega$. Then for ${\rm Re }(\lambda)>0$ the integral kernel of the Poisson transform $(g,k)\rightarrow p_\lambda(g,k):= a(g^{-1}k)^{-(\rho+\lambda)} \gamma(\kappa(g^{-1}k))T$ is a smooth function from $VM$ to $L^2(U,V(\gamma))\otimes V_{\tilde{\sigma}}$ (a much more detailed analysis of the Poisson kernel is given below in the proof of Lemma \ref{th43}). If $\phi\in {}^\Gamma C^{-\infty}( \Lambda ,V(\sigma_\lambda)) $, then $P \phi $ is $\Gamma$-invariant, and since ${\rm Re }(\lambda)>0$ it descends to a section in $L^2(Y,V_Y(\gamma))$. Moreover it is annihilated by $(-\Omega_G+c_\sigma+\lambda^2)$. Since $Y$ is complete $\Omega_G$ is essentially selfadjoint on the domain $C_c^\infty(Y,V_Y(\gamma))$. Its selfadjoint closure has the domain of definition $\{f\in L^2(Y,V_Y(\gamma))| \Omega_Gf\in L^2(Y,V_Y(\gamma))\}$. In particular, $\Omega_G$ can not have non-trivial eigenvectors in $L^2(Y,V_Y(\gamma))$ to eigenvalues with non-trivial imaginary part. Since ${\rm Im}(\lambda^2)\not=0$ we conclude that $P\phi =0$ and hence $\phi=0$ by the injectivity of the Poisson transform. \hspace*{\fill}$\Box$ \\[0.5cm]\noindent \section{Meromorphic continuation of $ext$}\label{extttttt} The extension $ext$ forms a holomorphic family of maps depending on $\lambda\in{\aaaa_\C^\ast}$ (we have omitted this dependence in order to simplify the notation) which is defined now for ${\rm Re }(\lambda)>\delta_\Gamma$. In the present section we consider the meromorphic continuation of $ext$ to (almost) all of ${\aaaa_\C^\ast}$. Our main tool is the scattering matrix which we introduce below. The scattering matrix for the trivial group $\Gamma=\{1\}$ is the Knapp-Stein intertwining operator of the corresponding principal series representation. We first recall basic properties of the Knapp-Stein intertwining operators. Then we define the scattering matrix using the extension and the Knapp-Stein intertwining operators. We simultaneously obtain the meromorphic continuations of the scattering matrix and the extension map. If $\sigma$ is a representation of $M$, then we define its Weyl-conjugate $\sigma^w$ by $\sigma^w(m):=\sigma( w^{-1} mw)$, where $w\in N_K(M)$ is a representative of the non-trivial element of the Weyl group $\cong {\bf Z}_2$ of $({\bf g},{\bf a})$. The Knapp-Stein intertwining operators form meromorphic families of $G$-equivariant operators (see \cite{knappstein71}) $$\hat{J}_{\sigma,\lambda}:C^{*}(\partial X,V(\sigma_\lambda)) \rightarrow C^{*}(\partial X ,V(\sigma^w_{-\lambda})),\quad *= -\infty,\infty\ .$$ Here $\:\hat{}\:$ indicates that $\hat{J}_{\sigma,\lambda}$ is unnormalized. In order to fix our conventions we give a definition of $\hat{J}_{\sigma,\lambda}$ as an integral operator acting on smooth functions for ${\rm Re }(\lambda)<0$. Its continuous extension to distributions is obtained by duality. For ${\rm Re }(\lambda)\ge 0$ it is defined by meromorphic continuation. Consider $f\in C^{\infty}(\partial X,V(\sigma_\lambda))$ as a function on $G$ with values in $V_{\sigma_\lambda}$ satisfying the usual invariance condition with respect to $P$. For ${\rm Re }(\lambda)<0$ the intertwining operator is defined by the convergent integral \begin{equation}\label{furunkel} (\hat{J}_{\sigma,\lambda}f)(g):=\int_{\bar{N}} f(g w\bar{n}) d\bar{n}\ . \end{equation} For all irreducible $\sigma\in\hat{M}$ we fix a minimal $K$-type (see \cite{knapp86}, Ch. XV for all that) of the principal series representation $C^{\infty}(\partial X,V(\sigma_\lambda))$. Let $c_{\sigma}(\lambda)$ be the value of $\hat{J}_{\sigma^w,-\lambda }$ on this minimal $K$-type. Then $c_{ \sigma}(\lambda)$ is a meromorphic function on ${\aaaa_\C^\ast}$ and we define the normalized intertwining operators by $$J_{ \sigma,\lambda}:= c_{ \sigma^w}^{-1}(-\lambda )\hat{J}_{\sigma,\lambda}\ .$$ Let $P_\sigma(\lambda):=c_\sigma(\lambda)^{-1}c_\sigma(-\lambda)^{-1}$ be the Plancherel density. Then the intertwining operators satisfy the following functional equation. \begin{equation}\label{spex} \hat{J}_{\sigma,\lambda}\circ \hat{J}_{\sigma^w,-\lambda}=\frac{1}{P_\sigma(\lambda)},\quad \quad J_{\sigma^{w },-\lambda }\circ J_{ \sigma,\lambda}={\mbox{\rm id}}\ .\end{equation} Our next goal is to show that the intertwining operators form a meromorphic family of operators in the sense defined in Section \ref{anaprep}. This is an easy application of the approach to the intertwining operators developed by Vogan-Wallach (see \cite{wallach92}, Ch. 10). The additional point we have to verify is that in the domain of convergence of (\ref{furunkel}) the operators $\hat{J}_{\sigma,\lambda}$ indeed form a continuous family. \begin{lem}\label{iny1} For ${\rm Re }(\lambda)<0$ the intertwining operators $$\hat{J}_{\sigma,\lambda}:C^{\infty}(\partial X,V(\sigma_\lambda)) \rightarrow C^{\infty}(\partial X ,V(\sigma^w_{-\lambda}))$$ form a holomorphic family of continuous operators. \end{lem} {\it Proof.$\:\:\:\:$} Let $X_i$, $i=1,\dots,\dim({\bf k})$, be an orthonormal base of ${\bf k}$. For any multiindex $r=(i_1,\dots,i_{\dim({\bf k})})$ we set $X_r=\prod_{l=1}^{\dim({\bf k})} X_l^{i_l}$, $|r|=\sum_{l=1}^{\dim({\bf k})} i_l$, and for $f\in C^{\infty}(K,V_{\sigma_\lambda})$ we define the seminorm $$\|f\|_r=\sup_{k\in K} |f(X_rk)|\ .$$ It is well known that the system $\{\|.\|_r\}$, $r$ running over all multiindices, defines the Fr\'echet topology of $C^{\infty}(K,V_{\sigma_\lambda})$ and by restriction the topology of $C^{\infty}(\partial X,V(\sigma_\lambda))$. We extend $f\in C^{\infty}(K,V_{\sigma_\lambda})$ to a function $f_\lambda$ on $G$ by setting $f_\lambda(kan):=f(k)a^{\lambda-\rho}$. Then we can define $$\hat{J}_{\sigma,\lambda} (f)(k)=\int_{\bar{N}} f_\lambda (k w\bar{n}) d\bar{n}\ .$$ For any $\lambda_0\in {\aaaa_\C^\ast}$ with ${\rm Re }(\lambda)<0$ and $\delta>0$ we can find an $\epsilon>0$ such that for $|\lambda-\lambda_0|<\epsilon $ $$ \int_{\bar{N}} |a(\bar{n})^{\lambda_0-\rho}-a(\bar{n})^{\lambda-\rho}| d\bar{n} <\delta\ .$$ We then have \begin{eqnarray*} \|\hat{J}_{\sigma, \lambda_0} f -\hat{J}_{\sigma,\lambda} f \|_r&=&\sup_{k\in K} \int_{\bar{N}} (f_{\lambda_0} (X_rk w\bar{n}) - (f_{\lambda } (X_rk w\bar{n}) ) d\bar{n}\\ &=& \sup_{k\in K} \int_{\bar{N}} f(X_rkw\kappa(\bar{n})) (a(\bar{n})^{\lambda_0-\rho}-a(\bar{n})^{\lambda-\rho})d\bar{n}\\ &\le& \|f\|_r \int_{\bar{N}} |a(\bar{n})^{\lambda_0-\rho}-a(\bar{n})^{\lambda-\rho}|d\bar{n}\\ &\le& \delta \|f\|_r \end{eqnarray*} This immediately implies that $\lambda\mapsto \hat{J}_{\sigma,\lambda}$ is a continuous family of operators on the space of smooth functions. The fact that the family $\hat{J}_{\sigma,\lambda}$, ${\rm Re }(\lambda)<0$, depends holomorphically on $\lambda$ is now easy to check (apply \cite{wallach92}, Lemma 10.1.3 and Fact \ref{holla}). \hspace*{\fill}$\Box$ \\[0.5cm]\noindent \begin{lem}\label{jjj} The family of intertwining operators $$\hat{J}_{\sigma,\lambda}:C^{\infty}(\partial X,V(\sigma_\lambda)) \rightarrow C^{\infty}(\partial X ,V(\sigma^w_{-\lambda}))$$ extends meromorphically to all of ${\aaaa_\C^\ast}$. \end{lem} {\it Proof.$\:\:\:\:$} We employ \cite{wallach92}, Thm. 10.1.5, which states that there are polynomial maps $b:{\aaaa_\C^\ast}\rightarrow {\bf C}$ and $D:{\aaaa_\C^\ast}\rightarrow {\cal U}({\bf g})^K$, such that \begin{equation}\label{shifty} b(\lambda)\hat{J}_{\sigma,\lambda}=\hat{J}_{\sigma, \lambda-4\rho}\circ \pi^{\sigma,\lambda-4\rho}(D(\lambda))\ . \end{equation} This formula requires some explanation. We identify $$C^{\infty}(\partial X, V(\sigma_\lambda))\cong C^\infty(K,V_\sigma)^M$$ canonically. Then all operators act on the same space $C^\infty(K,V_\sigma)^M$. If we know that $\hat{J}_{\sigma,\lambda}$ is meromorphic up to ${\rm Re }(\lambda)<\mu$, then we conclude that $$\hat{J}_{\sigma,\lambda}= \frac{1}{b(\lambda) } \hat{J}_{\sigma,\lambda-4\rho}\circ \pi^{\sigma,\lambda-4\rho}(D(\lambda))$$ is meromorphic up to ${\rm Re }(\lambda)<\mu+4\rho$. Thus the lemma follows from Lemma \ref{iny1}. \hspace*{\fill}$\Box$ \\[0.5cm]\noindent We call $\lambda\in {\aaaa_\C^\ast}$ integral if $$ 2 \frac{\langle \lambda,\alpha\rangle }{\langle\alpha,\alpha\rangle} \in{\bf Z}\quad \mbox{for some root $\alpha$ of $({\bf g},{\bf a})$}\ ,$$ and non-integral otherwise. The integral points form a discrete lattice in ${\bf a}^*$. It is known that the poles of $\hat{J}_{\sigma,\lambda}$ have at most order one and appear only at integral $\lambda$ (\cite{knappstein71}, Thm. 3 and Prop. 43). \begin{lem}\label{off} Let $\chi,\phi\in C^\infty(\partial X)$ such that ${\mbox{\rm supp}}(\phi)\cap {\mbox{\rm supp}}(\chi)=\emptyset$. Then $\chi\hat{J}_{\sigma,\lambda}\phi$ is a holomorphic family of smoothing operators. In particular, the residues of $\hat{J}_{\sigma,\lambda}$ are differential operators. \end{lem} {\it Proof.$\:\:\:\:$} Since ${\mbox{\rm supp}}(\phi)\cap{\mbox{\rm supp}}(\chi)=\emptyset$, there exists a compact set $V\subset \bar{N}$ such that $$\kappa({\mbox{\rm supp}}(\chi)M w (\bar{N}\setminus V)M \subset (\partial X\setminus {\mbox{\rm supp}}(\phi))M\ .$$ For ${\rm Re }(\lambda)<0$ and $f\in C^\infty(\partial X, V(\sigma_\lambda))$ we have (viewing $f$ as a function on $K$ with values in $V_{\sigma_\lambda})$ \begin{eqnarray*} (\chi \hat{J}_{\sigma,\lambda} \phi f)(k) &=& \int_{\bar{N}} \chi(k)f(\kappa(kw\bar{n}))\phi(\kappa(kw\bar{n})) a(\bar{n})^{ \lambda-\rho}d\bar{n}\\ &=& \int_{ V }\chi(k) f(\kappa(kw\bar{n})) \phi(\kappa(kw\bar{n})) a(\bar{n})^{ \lambda-\rho}d\bar{n}\ . \end{eqnarray*} The right-hand side of this equation extends to all of ${\aaaa_\C^\ast}$ and defines a holomorphic family of operators. This proves the first part of the lemma. It in particular implies that the residues of $\hat{J}_{\sigma,\lambda}$ are local operators. Hence the second assertion follows. \hspace*{\fill}$\Box$ \\[0.5cm]\noindent We need the following consequence of Lemma \ref{off}. Let $W\subset \partial X$ be a closed subset and let $${\cal G}_\lambda:=\{f\in C^{-\infty}(\partial X, V(\sigma_\lambda))| f_{|\partial X\setminus W}\in C^{\infty}(\partial X\setminus W, V(\sigma_\lambda))\}\ . $$ We equip ${\cal G}_\lambda$ with the weakest topology such that the maps ${\cal G}_\lambda\hookrightarrow C^{-\infty}(\partial X, V(\sigma_\lambda))$ and ${\cal G}_\lambda\rightarrow C^{\infty}(\partial X\setminus W, V(\sigma_\lambda))$ are continuous. Let $U\subset \bar{U}\subset \partial X\setminus W$ be open. \begin{lem}\label{dissmo} The composition $$res_U \circ \hat{J}_{\sigma,\lambda}:{\cal G}_\lambda\rightarrow C^\infty(U,V(\sigma^w_{-\lambda}))$$ is well-defined and depends meromorphically on $\lambda$. \end{lem} We introduce a notational convention concerning $\sigma$. Below $\sigma$ shall always denote a Weyl-invariant representation of $M$ which is either irreducible or of the form $\sigma^\prime\oplus \sigma^{\prime w}$ with $\sigma^\prime$ irreducible and not Weyl-invariant. In the latter case $c_\sigma(\lambda):=c_{\sigma^\prime}(\lambda)=c_{\sigma^{\prime w}}(\lambda)$, $P_\sigma(\lambda):=P_{\sigma^\prime}(\lambda)=P_{\sigma^{\prime w} }(\lambda)$. We omit the subscript $\sigma$ in the notation of the intertwining operators. We now turn to the definition of the (normalized) scattering matrix as a family of operators $$\hat{S}_\lambda \:( S_{\lambda})\: :C^{ *}(B,V_B(\sigma_\lambda))\rightarrow C^{*}(B,V_B(\sigma_{-\lambda })),\quad *=\infty,-\infty \ .$$ \begin{ddd}\label{scatdef} For ${\rm Re }(\lambda)>\delta_\Gamma$ we define \begin{equation}\label{scatde} \hat{S}_{ \lambda} :=res\circ \hat{J}_{\lambda}\circ ext\quad, \quad S_{ \lambda} :=res\circ J_{\lambda}\circ ext \ .\end{equation} \end{ddd} \begin{lem}\label{kkk} For ${\rm Re }(\lambda)>\delta_\Gamma$ the scattering matrix forms a meromorphic family of operators $$C^{\pm\infty}(B,V_B(\sigma_\lambda))\rightarrow C^{\pm\infty}(B,V_B(\sigma_{-\lambda }))\ .$$ If $\hat{S}_\lambda$ is singular and ${\rm Re }(\lambda)>\delta_\Gamma$, then $\lambda$ is integral and the residue of $\hat{S}_\lambda$ is a differential operator. \end{lem} {\it Proof.$\:\:\:\:$} The assertion for the scattering matrix acting on distributions follows from Lemma \ref{jjj}, Lemma \ref{off} and Fact \ref{comp}. The fact that the scattering matrix restricts to smooth sections follows from Lemma \ref{dissmo}.\hspace*{\fill}$\Box$ \\[0.5cm]\noindent \begin{lem}\label{lok} If ${\rm Re }(\lambda)>\delta_\Gamma$, then the adjoint $${}^tS_{ \lambda}: C^{\infty}(B,V_B(\tilde{\sigma}_\lambda))\rightarrow C^{\infty}(B,V_B(\tilde{\sigma}_{-\lambda }))$$ of $$S_{ \lambda}: C^{-\infty}(B,V_B(\sigma_\lambda))\rightarrow C^{-\infty}(B,V_B(\sigma_{-\lambda }))$$ coincides with the restriction of $S_\lambda$ to $C^{\infty}(B,V_B(\tilde{\sigma}_\lambda))$. \end{lem} {\it Proof.$\:\:\:\:$} We employ the fact that the corresponding relation holds for the intertwining operators (step (\ref{zx2}) below). Let $F\subset \Omega$ be a fundamental domain of $\Gamma$ and $\chi\in C_c^\infty(\Omega)$ be a cut-off function with $F\subset {\mbox{\rm supp}}(\chi)$ and $\sum_{g\in\Gamma} g^*\chi\equiv 1$ on $\Omega$. Let $\phi\in C^{\infty}(B,V_B(\sigma_\lambda))$, $f\in C^{\infty}(B,V_B(\tilde{\sigma}_\lambda))$, and consider $\phi$ as a distribution section. Then \begin{eqnarray} \langle \phi ,{}^tS_\lambda f\rangle &=&\langle S_\lambda\phi , f\rangle = \langle res\circ J_\lambda\circ ext \phi , f\rangle\nonumber\\ &=&\langle J_\lambda\circ ext \phi , \chi f \rangle = \langle ext \phi , {}^t J_\lambda \chi f \rangle \label{zx1}\\ &=&\langle ext \phi , J_\lambda \chi f \rangle\label{zx2}\\ &=&\langle \chi \phi , \sum_{g\in\Gamma } res_\Omega \circ \pi(g) J_\lambda \chi f\rangle = \sum_{g\in\Gamma } \langle \chi \phi , res_\Omega \circ \pi(g) J_\lambda \chi f\rangle\label{zx3}\\ &=& \sum_{g\in\Gamma } \langle \chi \phi , J_\lambda \pi(g) \chi f \rangle = \sum_{g\in\Gamma } \langle \pi(g) J_\lambda \chi \phi , \chi f\rangle\nonumber\\ &=& \langle \phi , S_\lambda f\rangle\ . \label{zx4} \end{eqnarray} Here in (\ref{zx1}) and (\ref{zx3}) we view $\chi f$ and $\chi \phi$ as sections over $\Omega$, respectively, and $\pi(g)$ is induced by the corresponding principal series representations. In order to obtain (\ref{zx4}) from the preceding line we do the transformations backwards with the roles of $\phi$ and $f$ interchanged. \hspace*{\fill}$\Box$ \\[0.5cm]\noindent \begin{lem}\label{pfun} If $|{\rm Re }(\lambda)|<- \delta_\Gamma$, then the scattering matrix satisfies the functional equation (viewed as an identity of meromorphic families of operators) $$S_{-\lambda}\circ S_\lambda = {\mbox{\rm id}}\ .$$ \end{lem} {\it Proof.$\:\:\:\:$} We employ Lemmas \ref{mainkor}, \ref{pokm}, and (\ref{spex}) in order to compute for ${\rm Im}(\lambda)\not= 0$, ${\rm Re }(\lambda)\not=0$, \begin{eqnarray*} S_{ -\lambda }\circ S_{ \lambda}&=&res\circ J_{ -\lambda }\circ ext\circ res \circ J_{ \lambda}\circ ext\\ &=&res\circ J_{ -\lambda }\circ J_{ \lambda}\circ ext \\ &=&res\circ ext\\ &=& {\mbox{\rm id}}\ . \end{eqnarray*} This identity now extends meromorphically to all of $\{|{\rm Re }(\lambda)|< -\delta_\Gamma\}$. \hspace*{\fill}$\Box$ \\[0.5cm]\noindent Now we start with the main topic of the present section, the meromorphic continuation of $ext$. We first invoke the meromorphic Fredholm theory Lemma \ref{merofred} in order to provide a meromorphic continuation of the scattering matrix to almost all of ${\aaaa_\C^\ast}$ under the condition $\delta_\Gamma<0$. We then use this continuation of the scattering matrix in order to define the meromorphic continuation of $ext$. Finally, if $X$ is not the Cayley hyperbolic plane, we drop the assumption $\delta_\Gamma<0$. To be more precise our method for the meromorphic continuations breaks down at a countable number of points. Therefore we introduce the set \begin{equation}\label{acagood}{\aaaa_\C^\ast}(\sigma):=\{\lambda\in{\aaaa_\C^\ast}\:|\: {\rm Re }(\lambda)>\delta_\Gamma\:\mbox{or}\: P_\sigma(\lambda) \hat{J}_{-\lambda}\:\mbox{is regular}\}\ .\end{equation} Not only the poles of $\hat{J}_{-\lambda}$, but also the poles of $P_\sigma(\lambda)$ are located at integral $\lambda$ (see e.g. the explicit formla (\ref{blanche}) given in the proof of Lemma \ref{casebycase}). Therefore, ${\aaaa_\C^\ast}(\sigma)$ contains all non-integral points, and many of the integral points, too. If $\delta_\Gamma<0$, then we show the meromorphic continuation of $ext$ and of the scattering matrix for $\lambda\in {\aaaa_\C^\ast}(\sigma)$. This result is employed in the proof of Proposition \ref{mystic}. If $\delta_\Gamma\ge 0$, then for simplicity we show the meromorphic continuation of $ext$ and the scattering matrix to the set of all non-integral $\lambda\in{\aaaa_\C^\ast}$, only. \begin{prop}\label{part1} The scattering matrix $$\hat{S}_\lambda :C^{\pm\infty}(B,V_B(\sigma_\lambda))\rightarrow C^{\pm\infty}(B,V_B(\sigma_{-\lambda }))$$ and the extension map $$ext: C^{-\infty}(B,V_B(\sigma_\lambda))\rightarrow {}^\Gamma C^{-\infty}(\partial X,V(\sigma_\lambda))$$ have meromorphic continuations to the set of all non-integral $\lambda\in{\aaaa_\C^\ast}$. In particular, we have \begin{equation}\label{forme} ext = J_{-\lambda} \circ ext \circ S_{\lambda}, \quad S_{-\lambda}\circ S_\lambda = {\mbox{\rm id}}\ .\end{equation} Moreover, $ext$ and $S_\lambda$ have at most finite-dimensional singularities at non-integral $\lambda$. \end{prop} {\it Proof.$\:\:\:\:$} We first assume that $\delta_\Gamma<0$. We construct the meromorphic continuation of $$S_\lambda:C^{\infty}(B,V_B(\sigma_\lambda))\rightarrow C^{\infty}(B,V_B(\sigma_{-\lambda }))\ ,$$ and then we extend this continuation to distributions by duality using Lemma \ref{lok}. The idea is to set $S_\lambda:=S_{-\lambda}^{-1}$ for ${\rm Re }(\lambda)<-\delta_\Gamma$ and to show that $S_{-\lambda}^{-1}$ forms a meromorphic family. Let $\{U_\alpha\}$ be a finite open covering of $B$ and let $\tilde{U}_\alpha$ be diffeomorphic lifts of $U_\alpha$. Choose a subordinated partition of unity $\phi_\alpha$. We view $\phi_\alpha$ as a smooth compactly supported function on $\tilde{U}_\alpha$. For $h\in\Gamma$ we set $\phi_\alpha^h(x):=\phi_\alpha(h^{-1}x)$. Let $1\in L\subset \Gamma$ be a finite subset. Then we define $\chi\in {}^\Gamma C^\infty(\Omega\times\Omega)$ by $$\chi(x,y):=\sum_{g\in\Gamma,h\in L,\alpha} \phi_\alpha(gx)\phi_\alpha^h(gy)\ .$$ Let $$\hat{J}^{diag}_{\lambda}:C^\infty(B, V_B(\sigma_\lambda))\rightarrow C^\infty(B, V_B(\sigma_{-\lambda}))$$ be the meromorphic family of operators obtained by multiplying the distributional kernel of $\hat{J}_{ \lambda}$ by $\chi$. If $f\in C^\infty(B, V_B(\sigma_\lambda))$, then $$(\hat{J}^{diag}_{ \lambda})f =\sum_{\alpha,h\in L} \phi_\alpha \hat{J}_\lambda(\phi^h_\alpha f)$$ using the canonical identifications. Below we shall employ the fact that $\hat{J}^{diag}_{ \lambda}$ depends on $L$. Let $$U:=\{\lambda\in{\aaaa_\C^\ast}\:|\:{\rm Re }(\lambda)>\delta_\Gamma,\: -\lambda\in{\aaaa_\C^\ast}(\sigma),\: \lambda\:\mbox{non-integral if}\:{\rm Re }(\lambda)\le 0\}\ .$$ Then $U$ is open and connected. For $\lambda\in U$ define \begin{equation}\label{klio}R(\lambda):=P_\sigma(\lambda) \hat{J}^{diag}_{-\lambda} \circ \hat{S}_\lambda - {\mbox{\rm id}}\ .\end{equation} The inverse of the unnormalized scattering matrix for $\lambda\in U$ should be given by \begin{equation}\label{finit}\hat{S}_{ \lambda}^{-1}=P_\sigma(\lambda) ({\mbox{\rm id}}+ R(\lambda))^{-1}\circ \hat{J}^{diag}_{ -\lambda }\ .\end{equation} It exists as a meromorphic family if $({\mbox{\rm id}}+ R(\lambda))^{-1}$ does. We want to apply the meromorphic Fredholm theory (Proposition \ref{merofred}) in order to invert ${\mbox{\rm id}} +R(\lambda)$ for $\lambda\in U$ and to conclude that $({\mbox{\rm id}}+R(\lambda))^{-1}$ is meromorphic. We check the assumption of Proposition \ref{merofred}. We choose a Hermitian metric on $V_B(\sigma_0)$ and a volume form on $\Omega$. The Hilbert space ${\cal H}$ of Proposition \ref{merofred} is $L^2(B,V_B(\sigma_0))$ defined using these choices. The Fr\'echet space ${\cal F}$ is just $C^\infty(B, V_B(\sigma_0))$. Implicitly, we identify the spaces $C^\infty(B, V_B(\sigma_\lambda))$ with $C^\infty(B, V_B(\sigma_0))$ using a trivialization of the holomorphic family of bundles $\{V_B(\sigma_\lambda)\}$, $\lambda\in{\aaaa_\C^\ast}$. We claim that $R(\lambda)$ is a holomorphic family of smoothing operators on $U$. If $\lambda\in U$, then $\hat{J}^{diag}_{-\lambda}$ as well as $$P_\sigma(\lambda)\hat{S}_\lambda=res\circ P_\sigma(\lambda)\hat{J}_\lambda\circ ext$$ are regular. Moreover, $R(\lambda)$ is smoothing by Lemma \ref{off}. Hence $R(\lambda)$ is indeed a holomorphic family of smoothing operators and this proves the claim. Next we show that if $L\subset \Gamma$ is sufficiently exhausting, then ${\mbox{\rm id}}+ R(\lambda)$ is injective for some $\lambda\in U$. Since ${\mbox{\rm id}}+R(\lambda)$ is Fredholm of index zero it is then invertible at this point. Here is one of the two instances where we assume $\delta_\Gamma<0$. We fix some non-integral $\lambda\in {\aaaa_\C^\ast}$ with $|{\rm Re }(\lambda)|<-\delta_\Gamma$. Define $\hat{J}^{off}_{-\lambda}:=res\circ\hat{J}_{-\lambda}-\hat{J}^{diag}_{-\lambda} \circ res$. By Lemma \ref{off} the composition $R(\lambda) = - P_\sigma(\lambda) \hat{J}^{off}_{-\lambda} \circ \hat{J}_{\lambda}\circ ext$ is a bounded operator on $C^k(B,V_B(\sigma_\lambda))$. The proof of Lemma \ref{anal1} shows that for ${\rm Re }(\lambda)>\delta_\Gamma$ the push down is a continuous map $\pi_\ast:C^0(\partial X,V(\tilde{\sigma}_{-\lambda}))\rightarrow C^0(B,V_B(\tilde{\sigma}_{-\lambda}))$. Thus the adjoint $ext$ restricts to a continuous map between the Banach spaces of measures of bounded variation $$ext: M_b(B,V_B(\sigma_\lambda))\rightarrow M_b(\partial X,V(\sigma_\lambda))\ .$$ The operator $\hat{J}_\lambda$ is a singular integral operator composed with a differential operator (as explained in Lemma \ref{jjj}). Thus there is a $k\in{\bf N}_0$ such that $$\hat{J}_\lambda:M_b(\partial X,V(\sigma_\lambda))\rightarrow C^{k}(\partial X,V(\sigma_{\lambda}))^\prime$$ is continuous. The scattering matrix $\hat{S}_\lambda$ extends to a continuous operator $$\hat{S}_\lambda:M_b(B,V_B(\sigma_\lambda))\rightarrow C^{k}(B,V_B(\sigma_{\lambda}))^\prime$$ and, by dualization and Lemma \ref{lok}, to a continuous map $$\hat{S}_\lambda:C^k(B,V_B(\sigma_\lambda))\rightarrow C^{0}(B,V_B(\sigma_{-\lambda})) \ .$$ \begin{lem}\label{converg1} If $L$ runs over an increasing sequence of subsets exhausting $\Gamma$, then $(\hat{J}^{diag}_{-\lambda}-\hat{S}_{-\lambda})\to 0$ in the sense of bounded operators from $C^{0}(B,V_B(\sigma_{-\lambda}))$ to $C^{k}(B,V_B(\sigma_{\lambda}))$. \end{lem} {\it Proof.$\:\:\:\:$} It follows from the estimates proved in Lemma \ref{anal1} that $(f\mapsto \sum_{h\in L } \phi^h_\alpha f)$ tends to $(f\mapsto ext(f))$ in the sense of bounded operators between the Banach spaces $C^0(B,V_B(\sigma_\lambda))$ and $M_b(\partial X, V(\sigma_\lambda ))$. We apply Lemma \ref{dissmo} letting $W$ be a compact neighbourhood of $\Lambda$ and $U\subset \Omega$ contain $\tilde{U}_\alpha$ for all $\alpha$. It implies that $$\sum_{h\in \Gamma\setminus L }\phi_\alpha \hat{J}_{-\lambda} \phi^h_\alpha\to 0$$ in the sense of bounded operators from $C^0(B,V_B(\sigma_{-\lambda}))$ to $C^{k}(\partial X,V(\sigma_\lambda))$ for all $\alpha$. The assertion of the lemma now follows. \hspace*{\fill}$\Box$ \\[0.5cm]\noindent By Lemmas \ref{converg1}, \ref{pfun} and Equation (\ref{spex}) the operator $R(\lambda)$ tends to zero in the sense of bounded operators on $C^k(B,V_B(\sigma_\lambda))$ when $L$ runs over an increasing sequence of subsets exausting $\Gamma$. Thus if $L$ is large enough, then ${\mbox{\rm id}}+R(\lambda)$ is injective. We now have verified the assumptions of Proposition \ref{merofred}. We conclude that the family $\hat{S}_{\lambda}^{-1}$ is meromorphic for $\lambda\in U$. If $\hat{S}_{\lambda}^{-1}$ has a singularity and $\lambda$ is non-integral, then this singularity is finite-dimensional. Here is the second and main instance, where we need the assumption $\delta_\Gamma<0$. Namely, it implies that $$\{\lambda\in{\aaaa_\C^\ast}\:|\:-\lambda\in U\}\cup \{\lambda\in{\aaaa_\C^\ast}\:|\:{\rm Re }(\lambda)>\delta_\Gamma\}={\aaaa_\C^\ast}(\sigma)\ .$$ Furthermore, by Lemma \ref{pfun} we have $S_\lambda=S_{-\lambda}^{-1}$ on $\{\lambda\in{\aaaa_\C^\ast}\:|\:-\lambda\in U\}\cap \{\lambda\in{\aaaa_\C^\ast}\:|\:{\rm Re }(\lambda)>\delta_\Gamma\}$. Thus, setting $S_\lambda:=S_{-\lambda}^{-1}$ for $-\lambda\in U$ we obtain a well-defined continuation of $S_\lambda$ to all of ${\aaaa_\C^\ast}(\sigma)$. By duality this continuation extends to distributions still having the same finite-dimensional singularities at non-integral points. It remains to consider the extension map. We employ the scattering matrix in order to define for $\lambda\in{\aaaa_\C^\ast}(\sigma)$, ${\rm Re }(\lambda)<-\delta_\Gamma$ $$ ext_1 := J_{-\lambda} \circ ext \circ S_{\lambda} \ .$$ We claim that $ext=ext_1$. In fact since $res$ is injective on an open subset of $\{|{\rm Re }(\lambda)| < -\delta_\Gamma\}$, the computation \begin{eqnarray*} res\circ ext_1&=& res\circ J_{-\lambda} \circ ext \circ S_\lambda\\ &=& S_{-\lambda}\circ S_\lambda \\ &=&{\mbox{\rm id}} \end{eqnarray*} implies the claim. We now have constructed a meromorphic continuation of $ext$ to all of ${\aaaa_\C^\ast}(\sigma)$. The relation (\ref{forme}) between the scattering matrix and $ext$ follows by meromorphic continuation. This equation also implies that $ext$ has at most finite-dimensional singularities. We have finished the proof of Proposition \ref{part1} assuming $\delta_\Gamma<0$. The identities $$\hat{S}_\lambda =res\circ \hat{J}_\lambda \circ ext, \quad S_\lambda\circ S_{-\lambda}={\mbox{\rm id}} $$ extend to all of ${\aaaa_\C^\ast}(\sigma)$ by meromorphic continuation. We now show how to drop the assumption $\delta_\Gamma<0$ using the embedding trick and Assumption \ref{caly}. If $X$ is the Cayley hyperbolic plane, then by assumption $\delta_\Gamma<0$ and the proposition is already proved. Thus we can assume that $X$ belongs to a series of rank-one symmetric spaces. Let $\dots \subset G^n\subset G^{n+1}\subset \dots$ be the corresponding sequence of real, semisimple, linear Lie groups inducing embeddings of the corresponding Iwasawa constituents $K^n\subset K^{n+1}$, $N^n \stackrel{\scriptstyle \subset}{ \scriptstyle \not=} N^{n+1}$, $ M^n\subset M^{n+1}$ such that $A=A^n = A^{n+1}$. Then we have totally geodesic embeddings of the symmetric spaces $X^n\subset X^{n+1}$ inducing embeddings of their boundaries $\partial X^n\subset \partial X^{n+1}$. If $\Gamma\subset G^n$ satisfies \ref{asss} then it keeps satisfying \ref{asss} when viewed as a subgroup of $G^{n+1}$. We obtain embeddings $\Omega^n\subset \Omega^{n+1}$ inducing $B^n\subset B^{n+1}$ while the limit set $ \Lambda^n$ is identified with $ \Lambda^{n+1}$. Let $\rho^n(H)=\frac{1}{2}{\mbox{\rm tr}}({\mbox{\rm ad}}(H)_{|{\bf n}^n})$, $H\in{\bf a}$. The exponent of $\Gamma$ now depends on $n$ and is denoted by $\delta_\Gamma^n$. We have the relation $\delta_\Gamma^{n+1}=\delta_\Gamma^n-\rho^{n+1}+\rho^n$. Thus $\delta_\Gamma^{n+m}\to -\infty$ as $m\to 0$ and hence taking $m$ large enough we can satisfy $\delta_\Gamma^{n+m}<0$. The aim of the following discussion is to show how the meromorphic continuation $ext^{n+1}$ leads to the continuation of $ext^n$. Let $\sigma^{n+1}$ be a Weyl-invariant representation of $M^{n+1}$. Then it restricts to a Weyl-invariant representation of $M^n$. For any given finite-dimensional representation $\sigma^n$ of $M^n$ we can find a Weyl-invariant representation $\sigma^{n+1}$ of $M^{n+1}$ such that $\sigma^{n+1}_{|M^n}$ contains $\sigma^n$ as a subrepresentation. The representation $\sigma^{n+1}_\lambda$ of $P^{n+1}$ restricts to the representation $(\sigma^{n+1}_{|P^n})_{\lambda+\rho^n-\rho^{n+1}}$ of $P^n$. This induces an isomorphism of bundles $$V_{B^{n+1}}(\sigma^{n+1}_\lambda)_{|B^n}=V_{B^n}((\sigma^{n+1}_{|P^n})_{\lambda+\rho^n-\rho^{n+1}})\ .$$ We will omit the subscript ${}_{|P^n}$ and the superscript ${}^{n+1}$ of $\sigma$ in the following discussion. We obtain a push forward of distributions $$i_*:C^{-\infty}(B^n,V_{B^n}(\sigma_\lambda))\rightarrow C^{-\infty}(B^{n+1},V_{B^{n+1}}(\sigma_{\lambda+\rho^n-\rho^{n+1}}))\ .$$ For $\phi\in C^{-\infty}(B^n,V_{B^n}(\sigma_\lambda))$ the push forward $i_\ast(\phi)$ has support in $B^n\subset B^{n+1}$. Then also ${\mbox{\rm supp}}(ext(\phi))\subset \partial X^n$. We try to define a pull back $ext^n(\phi):=i^*\circ ext^{n+1}\circ i_*(\phi)$ as follows. For $f\in C^\infty(\partial X^{n},V(\tilde{\sigma}_{-\lambda}))$ let $\tilde{f}\in C^\infty(\partial X^{n+1},V(\tilde{\sigma}_{-\lambda-\rho^n+\rho^{n+1}}))$ be an arbitrary extension. Then we set $$\langle ext^n(\phi),f\rangle := \langle ext^{n+1}\circ i_*(\phi_\lambda),\tilde{f}\rangle\ .$$ \begin{lem}\label{well} $ext^n$ is well defined. \end{lem} We must show that this definition does not depend on the choice of the extension of $\tilde{f}$. It is sufficient to show that if $h\in C^\infty(\partial X^{n+1},V(\tilde{\sigma}_{-\lambda-\rho^n+\rho^{n+1}}))$ vanishes on $\partial X^{n}$, then $\langle ext^{n+1}\circ i_*(\phi_\lambda),h\rangle=0$. Embed $\phi,h$ into holomorphic families $\phi_\mu, h_\mu$ , $\phi_\mu\in C^{-\infty}(B^n,V_{B^n}(\sigma_\mu))$, $h_\mu\in C^\infty(\partial X^{n+1},V(\tilde{\sigma}_{-\mu-\rho^n+\rho^{n+1}}))$ such that $(h_\mu)_{|\partial X^{n} }=0$. Then for ${\rm Re }(\mu)$ large enough\linebreak[4] $\pi^{n+1}_*(h_{-\mu})_{|B^n}=0$ and thus $$\langle ext^{n+1}\circ i_*(\phi_\mu),h_{-\mu} \rangle= \langle i_*(\phi_\mu),\pi^{n+1}_*h_{-\mu}\rangle=0\ .$$ By meromorphic continuation this identity holds for all $\mu$, in particular at $\mu=\lambda$. \hspace*{\fill}$\Box$ \\[0.5cm]\noindent If $\lambda\in{\aaaa_\C^\ast}$ is non-integral, then so is $\lambda+\rho^n-\rho^{n+1}$. We deduce the properties of $ext^{n}$ from the corresponding properties of $ext^{n+1}$. In particular, $ext^n$ is continuous, meromorphic, and has at most finite-dimensional singularities at $\lambda$ if $ext^{n+1}$ has these properties at $\lambda+\rho^n-\rho^{n+1}$. We define the meromorphic continuation of the scattering matrix by (\ref{scatde}). Then it is easy to see that the scattering matrix has the properties as asserted. This finishes the proof of Proposition \ref{part1}. \hspace*{\fill}$\Box$ \\[0.5cm]\noindent \begin{lem}\label{lead} If $ext$ is meromorphic at $\lambda\in {\aaaa_\C^\ast}$ and ${}^\Gamma C^{-\infty}(\Lambda,V(\sigma_\lambda))=0$, then $ext$\ is regular at $\lambda$. \end{lem} {\it Proof.$\:\:\:\:$} If $ext$ is meromorphic, then for any holomorphic family $\mu\to\phi_\mu\in C^{-\infty}(B,V_B(\sigma_\mu))$ the leading singular part of $ext(\phi_\mu)$ at $\mu=\lambda$ belongs to ${}^\Gamma C^{-\infty}(\Lambda,V(\sigma_\lambda))$. In fact $res\circ ext(\phi_\mu)=\phi_\mu$ has no singularity. If ${}^\Gamma C^{-\infty}(\Lambda,V(\sigma_\lambda))=0$, then $ext(\phi_\mu)$ is regular at $\mu=\lambda$ for any holomorphic family $\mu\to \phi_\mu$.\hspace*{\fill}$\Box$ \\[0.5cm]\noindent In Lemma \ref{ghu} below we consider the converse of Lemma \ref{lead} for non-integral $\lambda\in{\aaaa_\C^\ast}$ with ${\rm Re }(\lambda)>0$. \begin{lem}\label{firtu} If ${\rm Re }(\lambda)>0$ and $ext:C^{-\infty}(B,V_B(\sigma_\mu))\rightarrow {}^\Gamma C^{-\infty}(\partial X,V(\sigma_\mu))$ is meromorphic at $\mu=\lambda$, then the order of a singularity of $ext$ at $\lambda$ is at most $1$. \end{lem} {\it Proof.$\:\:\:\:$} Let $\gamma\in \hat{K}$ be such that there exists an injective $T\in {\mbox{\rm Hom}}_M(V_\sigma,V_\gamma)$. We also can and will require that the Poisson transform $P_\lambda^T=:P$ is injective (e.g. by taking $\gamma$ to be the minimal $K$-type of the principal series representation $\pi^{\sigma,\lambda}$). Let $f_\mu\in C^\infty(B,V_B(\sigma_\mu))$, $\mu\in {\aaaa_\C^\ast}$, be a holomorphic family such that $ext(f_\mu)$ has a pole of order $n\ge 1$ at $\mu=\lambda$, ${\rm Re }(\lambda)>0$. We assume that $n\ge 2$ and argue by contradiction. Let $0\not=\phi \in {}^\Gamma C^{-\infty}(\Lambda,V(\sigma_\lambda))$ be the leading singular part of $ext(f_\mu)$ at $\mu=\lambda$. Then $(\lambda^2-\mu^2)^{n-1} ext(f_\mu)$ and hence $(\lambda^2-\mu^2)^{n-1} P_\mu^T ext(f_\mu)$ have first-order poles, the latter with residue $-(2\lambda)^{n-1} P\phi $. Since $res_\Omega\circ ext(f_\mu)$ is smooth if $kM\in\Omega M$ we have $P^T_\mu\circ ext(f_\mu) (ka)=O(a^{\mu-\rho})$. Moreover $P\phi (ka)=O(a^{-\lambda-\rho})$ and both estimates hold uniformly for $kM$ in compact subsets of $\Omega$, $\mu$ near $\lambda$, and large $a\in A_+$. This justifies the following computation using partial integration: \begin{eqnarray*} \infty&=&\lim_{\mu\to\lambda,\: {\rm Re }(\mu)<{\rm Re }(\lambda)} \langle (\lambda^2-\mu^2)^{n-1} P_\mu^T ext(f_\mu), P\phi \rangle_{L^2(Y)}\\ &=&\lim_{\mu\to\lambda,\: {\rm Re }(\mu)<{\rm Re }(\lambda)} \langle (-\Omega_G+c_\sigma+\lambda^2)^{n-1} P_\mu^T ext(f_\mu), P\phi \rangle_{L^2(Y)}\\ &=& 0 \end{eqnarray*} This is a contradiction and thus $n=1$. \hspace*{\fill}$\Box$ \\[0.5cm]\noindent Since $\sigma$ is an unitary representation of $M$, we have for $\lambda\in \imath{\bf a}^*$ a positive conjugate linear pairing $V_{\sigma_\lambda}\otimes V_{\sigma_\lambda} \rightarrow V_{1_{-\rho}}$ and hence a natural $L^2$-scalar product $C^\infty(B,V_B(\sigma_\lambda))\otimes C^\infty(B,V_B(\sigma_\lambda))\rightarrow {\bf C}$. Let $L^2(B,V_B(\sigma_\lambda))$ be associated Hilbert space. Using Lemma \ref{lok} we see that the adjoint $S^*_\lambda$ with respect to this Hilbert space structure is just $S_{-\lambda}$. \begin{lem}\label{unitary} If ${\rm Re }(\lambda)=0$, $\lambda\not=0$, then $S_\lambda$ is regular and unitary. \end{lem} {\it Proof.$\:\:\:\:$} The scattering matrix $S_\lambda$ is meromorphic at non-zero imaginary points $\lambda$. Let now $\lambda$ be imaginary, $S_{\pm\lambda}$ be regular and $f\in C^\infty(B,V_B(\sigma_\lambda))$. Then by the functional equation (\ref{forme}) $$\|S_\lambda f\|_{L^2(B,V_B(\sigma_\lambda))}^2=\langle S_{-\lambda}\circ S_\lambda f, f \rangle_{L^2(B,V_B(\sigma_\lambda))} = \|f\|^2_{L^2(B,V_B(\sigma_\lambda))}\ .$$ This equation remains valid at all non-zero imaginary points. Hence, $S_\lambda$ is regular and unitary there. \hspace*{\fill}$\Box$ \\[0.5cm]\noindent \section{Invariant distributions on the limit set}\label{invvv} In present section we study the space ${}^\Gamma C^{-\infty}(\Lambda,V(\sigma_\lambda))$ of invariant distributions which are supported on the limit set, mainly for ${\rm Re }(\lambda)\ge 0$. We show that nontrivial distributions of this kind can only exist for a countable set of parameters $\lambda\ge 0$ with possibly finitely many accumulation points. In particular, we show that ${}^\Gamma C^{-\infty}(\Lambda,V(\sigma_\lambda))=0$ if $ext$ is regular at $\lambda$, ${\rm Re }(\lambda)\ge 0$, and $\chi_{\mu_\sigma+\rho_m-\lambda}$ is non-integral. In the course of this paper we will prove the finiteness of the discrete spectrum of ${\cal Z}$ on $L^2(Y,V_Y(\gamma))$, ore equivalently, that nontrivial invariant distributions with support on the limit set (with ${\rm Re }(\lambda)\ge 0$) can in fact only exist for a finite set of parameters $\lambda\ge 0$. The proof of the finiteness of the point spectrum is essentially based on the spectral comparison Proposition \ref{esspec}. But this proposition is not applicable in order to exclude that eigenvalues accumulate at the boundary of the continuous spectrum. Here we will employ Corollary \ref{nahenull} instead, and it is important to show that $ext$ is meromorphic at $\lambda=0$ for certain $\sigma$. First we show a variant of Green's formula. We need nice cut-off functions which exist by the following lemma. \begin{lem}\label{lll} There exists a cut-off function $\chi$ such that \begin{enumerate} \item $\chi> 0$ on a fundamental domain $F\subset X$, \item ${\mbox{\rm supp}}(\chi)\subset \cup_{g\in L} gF$ for some finite subset $L\subset \Gamma$, \item $\sum_{g\in\Gamma}g^*\chi = 1$, \item $\sup_{k\in ({\rm clo}(F)\cap\Omega)M ,\: a\in A_+} a\:|\nabla^i\chi(ka)|<\infty$, $i\in{\bf N}$, \item the function $\chi_\infty$ on $K$ defined by $\chi_\infty(k):=\lim_{a\to\infty}\chi(ka)\chi_\infty(k)$ is smooth. \end{enumerate} \end{lem} {\it Proof.$\:\:\:\:$} Let $W\subset \Omega$ be compact such that ${\rm clo}(F)\cap \Omega\subset W$. Let $\psi\in C^\infty(\partial X)$ be a cut-off function such that $\psi_{|{\rm clo}(F)\cap\Omega}=1$ and $\psi_{|\partial X\setminus W}=0$. Let $B_R\subset X$, $R\in{\bf R}$ be the $R$-ball in $X$ centered at the origin and choose $R>1$ so large that $F\subset B_R\cup WA_+$. Let $\sigma\in C^\infty(A_+)$ be a cut-off function such that $\sigma(r)=1$ for $r>1$ and $\sigma(r)=0$ for $r<1/2$. Finally let $\phi\in C_c^{\infty}(X)$ be a cut-off function such that $\phi_{|B_R}=1$ and $\phi_{|\partial X\setminus B_{R+1}}=0$. Then we set $\tilde{\chi}(ka):=\phi(ka) + \psi(k)\sigma(a)$, $k\in K$, $a\in A_+$, $ka\in X$. If we define $$\chi:=\frac{\tilde{\chi}}{\sum_{g\in \Gamma }g^*\tilde{\chi}}\ ,$$ then $\chi$ obviously satisfies (1),(2), (3), and (5). It remains to verify (4). Note that by construction of $\tilde{\chi}$ for all $l\in {\bf N}$ there exists a constant $C<\infty$ such that for all $k\in K$, $a\in A_+$ $$|\nabla^l \tilde{\chi}(ka)| < C a^{-1}\ .$$ Hence for any finite subset $L\subset \Gamma$ and $l\in{\bf N}$ there exists a constant $C<\infty$ such that for all $k\in K$, $g\in L$, $a\in A_+$ $$|\nabla^l g^\ast \tilde{\chi}(ka)| < C a^{-1}\ .$$ This implies (4). \hspace*{\fill}$\Box$ \\[0.5cm]\noindent Now we consider the variant of Green's formula. Let $\phi\in {}^\Gamma C^{-\infty}(\Lambda,V(\sigma_\lambda))$. If ${\rm Re }(\lambda)\ge 0$, then $res\circ \hat{J}_\lambda(\phi)\in C^{\infty}(B,V_B(\sigma_{-\lambda}))$ is well defined even if $\hat{J}_\mu$ has a pole at $\mu=\lambda$. In the latter case the residue of $\hat{J}_\mu$ at $\mu=\lambda$ is a differential operator $D_\lambda$ (see Lemma \ref{off}) and $res\circ D_\lambda(\phi)=0$. \begin{prop}\label{green} Assume that ${\rm Re }(\lambda)\ge 0$ and $\lambda\not=0$. If $\phi\in {}^\Gamma C^{-\infty}(\Lambda,V(\sigma_\lambda))$ and $f\in {}^\Gamma C^{-\infty}(\partial X,V(\tilde{\sigma}_\lambda))$ is such that $f_{|\Omega}$ is smooth, then $$\langle res\circ J_\lambda (\phi),res(f)\rangle =0\ .$$ \end{prop} {\it Proof.$\:\:\:\:$} Let $(\gamma,V_\gamma) \in \hat{K}$ be a minimal $K$-type of the principal series representation of $G$ on $C^{\infty}(\partial X, V(\sigma_\lambda))$. Then there is an injective $T\in {\mbox{\rm Hom}}_M(V_\sigma,V_\gamma)$. For ${\rm Re }(\lambda)>0$ we define the endomorphism valued function \begin{equation}\label{cgamma} c_{\gamma}(\lambda):=\int_{\bar{N}} a(\bar{n})^{-(\lambda+\rho)} \gamma(\kappa(\bar{n}))d\bar{n}\in {\mbox{\rm End}}_M(V_\gamma)\ .\end{equation} The function $c_{\gamma}(\lambda)$ extends meromorphically to all of ${\aaaa_\C^\ast}$. If $c_\gamma(\lambda)$ is singular, then $\lambda$ is integral. We choose $\tilde{T}\in {\mbox{\rm Hom}}_M(V_{\tilde{\sigma}},V_{\tilde{\gamma}})$ such that $\langle \gamma(w) T v, c_\gamma(\lambda) \tilde{T}u\rangle =\langle v,u\rangle$ for all $v\in V_\sigma$, $u\in V_{\tilde{\sigma}}$. This is possible since $\gamma$ is a minimal $K$-type and hence $c_\gamma(\lambda)$ is invertible for ${\rm Re }(\lambda)\ge 0$, $\lambda\not=0$. Let $A=-\Omega_G+c_\sigma+\lambda^2$, $P=P_\mu$ be the Poisson transform (associated to $T$ or $\tilde{T}$, respectively), and $\chi$ be the cut-off function constructed in Lemma \ref{lll}. Note that $A=-\nabla^*\nabla + {\cal R}$ for some selfadjoint endomorphism ${\cal R}$ of $V(\gamma)$, where $-\nabla^*\nabla=\Delta$ is the Bochner Laplacian associated to the invariant connection $\nabla$ of $V(\gamma)$. By $B_R$ we denote the metric $R$-ball centered at the origin of $X$. The following is an application of Green's formula: \begin{eqnarray} 0&=&\langle \chi A P\phi,Pf\rangle_{L^2(B_R)}- \langle \chi P\phi,A Pf \rangle_{L^2(B_R)}\label{limiz}\\ &=& \langle A \chi P\phi,Pf\rangle_{L^2(B_R)}- \langle \chi P\phi,A Pf \rangle_{L^2(B_R)} - \langle [ A, \chi] P\phi,Pf\rangle_{L^2(B_R)}\nonumber \\ &=& - \langle \nabla_n \chi P\phi,Pf\rangle_{L^2(\partial B_R)}+ \langle \chi P\phi,\nabla_n Pf \rangle_{L^2(\partial B_R)} - \langle [ A, \chi] P\phi,Pf\rangle_{L^2(B_R)}\ , \nonumber \end{eqnarray} where $n$ is the exterior unit normal vector field at $\partial B_R$. For the following discussion we distinguish between the two cases ${\rm Re }(\lambda)>0$ and ${\rm Re }(\lambda)=0$, $\lambda\not=0$. We first consider the case ${\rm Re }(\lambda)>0$. We employ the following asymptotic behaviour of the Poisson transforms of $f$ and $\psi$. For $kM\in \Omega$ we have \begin{eqnarray} P(f)(ka)&=&a^{\lambda-\rho} c_\gamma(\lambda) \tilde{T} f(k)+O(a^{\lambda-\rho-\epsilon})\label{k0o}\\ P(\phi)(ka)&=&a^{-\lambda-\rho} \gamma(w)T (\hat{J}_\lambda \phi)(k) + O(a^{-\lambda-\rho-\epsilon})\label{k0o1} \end{eqnarray} where $\epsilon>0$. While (\ref{k0o}) follows from the fact that $f_{|\Omega}$ is smooth, (\ref{k0o1}) is shown in Lemma \ref{eee}. The estimate can be differentiated with respect to $a$ and holds locally uniformly on $\Omega$. By property (4) of $\chi$ we see that $\langle [ A, \chi] P\phi,Pf\rangle$ is integrable and by property (3) and the $\Gamma$-invariance of $f$ and $\phi$ we have $\langle [ A, \chi] P\phi,Pf\rangle_{L^2(X)}=0$. Taking the limit $R\to\infty$ in (\ref{limiz}) we obtain \begin{eqnarray*} 0&=& (\lambda+\rho)\int_{\partial X} \chi_\infty(k) \langle\gamma(w)T (\hat{J}_\lambda \phi )(k), c_\gamma(\lambda)\tilde{T} f(k)\rangle \\ &&+(\lambda-\rho)\int_{\partial X} \chi_\infty(k) \langle\gamma(w)T (\hat{J}_\lambda \phi )(k),c_\gamma(\lambda)\tilde{T}f(k)\rangle \\ &=&2\lambda \int_{\partial X} \chi_\infty(k) \langle (\hat{J}_\lambda \phi )(k),f(k)\rangle \\ &=&2\lambda \langle res\circ\hat{J}_\lambda (\phi),res( f)\rangle \ . \end{eqnarray*} This is the assertion of the proposition for ${\rm Re }(\lambda)>0$. Now we discuss the case ${\rm Re }(\lambda)=0$ and $\lambda\not=0$. In this case we have the following asymptotic behaviour $$P(f)(ka) = a^{\lambda-\rho} c_\gamma(\lambda) \tilde{T} f(k)+ a^{-\lambda-\rho}\gamma(w) \tilde{T} \hat{J}_\lambda f(k) + O(a^{ -\rho-\epsilon})\ .$$ Instead of taking the limit $R\to\infty$ in (\ref{limiz}) we apply $\lim_{r\to\infty}\frac{1}{r}\int_0^r dR$. Again we have $$\lim_{r\to\infty}\frac{1}{r}\int_0^r \langle [ A, \chi] P\phi,Pf\rangle_{L^2(B_R)} dR=0\ . $$ Moreover, the asymptotic term $a^{-\lambda-\rho} \tilde{T} \hat{J}_\lambda f(k)$ does not contribute to the limit because of $$-2\lambda \lim_{r\to\infty}\frac{1}{r}\int_0^r R^{-2\lambda} \langle \chi(.R) \gamma(w)T \hat{J}_\lambda \phi ,\gamma(w) \tilde{T} \hat{J}_\lambda f\rangle_{L^2(\partial X)} dR=0\ .$$ The contribution of the term $a^{\lambda-\rho} c_\gamma(\lambda) \tilde{T} f(k)$ leads to $$0=2\lambda \langle res\circ\hat{J}_\lambda (\phi),res( f)\rangle$$ as in the case ${\rm Re }(\lambda)>0$. \hspace*{\fill}$\Box$ \\[0.5cm]\noindent By the Harish-Chandra isomorphism characters of ${\cal Z}$ are parametrized by elements of ${\bf h}_{\bf C}^*/W$. A character $\chi_\lambda$, $\lambda\in {\bf h}_{\bf C}^*$, is called integral, if $$2\frac{\langle \lambda,\alpha\rangle}{\langle\alpha,\alpha\rangle}\in{\bf Z}$$ for all roots $\alpha$ of $({\bf g},{\bf h})$. \begin{lem}\label{th43} Let $\lambda\in {\aaaa_\C^\ast}$ be such that \begin{itemize} \item ${\rm Re }(\lambda)\ge 0$ and $\chi_{\mu_\sigma+\rho_m-\lambda}$ is a non-integral character of ${\cal Z}$ or \item ${\rm Re }(\lambda)<0$ and $\lambda$ is non-integral. \end{itemize} Let and $U\subset \partial X$ be open. If $\phi\in C^{-\infty}(\partial X, V(\sigma_\lambda))$ satisfies $\phi_{|U}= 0$, then $(\hat{J}_\lambda\phi)_{|U}= 0$ implies $\phi=0$. \end{lem} Before turning to the proof note that Lemma \ref{off} implies that $(\hat{J}_\lambda\phi)_{|U}$ is well-defined even if $\hat{J}_\lambda$ has a pole.\\[0.5cm]\noindent {\it Proof.$\:\:\:\:$} We modify an argument given by van den Ban-Schlichtkrull \cite{vandenbanschlichtkrull89} for the case $\sigma=1$. If $\bar{N}$ is two-step nilpotent, then $\bar{{\bf n}}=\bar{{\bf n}}_1\oplus \bar{{\bf n}}_2$, $[\bar{{\bf n}}_1,\bar{{\bf n}}_1]=\bar{{\bf n}}_2$, where $\bar{{\bf n}}_1$ corresponds to the negative of the shorter root $\alpha_1\in{\bf a}^*_+$. \begin{lem}\label{eee} Let $\lambda\in{\aaaa_\C^\ast}$, $\gamma$ be a representation of $K$, and $\phi\in C^{-\infty}(\partial X,V(\sigma_\lambda))$ satisfy $M \not\in {\mbox{\rm supp}}(\phi)$. Then for $T\in {\mbox{\rm Hom}}(V_\sigma,V_\gamma)$, $k\in K\setminus{\mbox{\rm supp}}(\phi)M$ the Poisson transform $P\phi:=P^T_\lambda\phi$ has an asymptotic expansion as $a\to\infty$ of the form \begin{equation}\label{epan} (P\phi)(ka)=a^{-(\lambda+\rho)}\gamma(w)T(\hat{J}_\lambda\phi)(k) +\sum_{n\ge 1} a^{-(\lambda+\rho)-n\alpha_1}\psi_n(k)\ . \end{equation} Here $\psi_n$ are smooth functions on $K\setminus{\mbox{\rm supp}}(\phi)M$. The expansion converges uniformly for $a>>0$ and $k$ in compact subsets of $K\setminus {\mbox{\rm supp}}(\phi)M$. \end{lem} {\it Proof.$\:\:\:\:$} In the following computation we write the pairing of a distribution with a smooth function as an integral. \begin{eqnarray} (P\phi)(ka)&=&\int_K a(a^{-1}k^{-1}h)^{-(\lambda+\rho)}\gamma(\kappa(a^{-1}k^{-1}h)) T \phi(h) dh\nonumber\\ &=& \int_{\bar{N}} a( a^{-1}w\kappa(\bar{n}))^{-(\lambda+\rho)} \gamma(\kappa( a^{-1}w\kappa(\bar{n}))) T\phi(kw\kappa(\bar{n})) a(\bar{n})^{-2\rho} d\bar{n} \nonumber\\ &=&\int_{\bar{N}} a(a \bar{n} a^{-1})^{-(\lambda+\rho)} a^{-(\lambda+\rho)} a(\bar{n})^{\lambda+\rho} \gamma(w) \gamma(\kappa(a\bar{n}a^{-1})) T\phi(kw\kappa(\bar{n})) a(\bar{n})^{-2\rho} d\bar{n} \nonumber\\ &=&a^{-(\lambda+\rho)}\gamma(w) \int_{\bar{N}} a(\bar{n})^{\lambda-\rho} a(a\bar{n}a^{-1})^{-(\lambda+\rho)} \gamma(\kappa(a\bar{n}a^{-1})) T \phi(kw\kappa(\bar{n})) d\bar{n}\label{tyh}\ . \end{eqnarray} For $z\in {\bf R}^+$ define $a_z\in A$ through $z=a_z^{-\alpha_1}$. We consider the map $\Phi:(0,\infty)\times \bar{N}\ni (z,\bar{n})\mapsto a_z\bar{n}a_z^{-1}\in \bar{N}$ which is can also be written as $$\Phi(z,exp(X+Y)):= exp(zX+z^2Y), \quad X\in{\bf n}_1, Y\in{\bf n}_2\ .$$ Thus $\Phi$ and hence $ (z,\bar{n})\mapsto a(a_z\bar{n}a_z^{-1})^{-(\lambda+\rho)} \gamma(\kappa(a_z\bar{n}a_z^{-1}))$ extend analytically to ${\bf R}\times\bar{N}$. Taking the Taylor expansion with respect to $z$ at $z=0$ we obtain $$a(a_z\bar{n}a_z^{-1})^{-(\lambda+\rho)} \gamma(\kappa(a_z\bar{n}a_z^{-1}))={\mbox{\rm id}} + \sum_{n\ge 1}A_n(\bar{n}) z^n\ .$$ Here $A_n:\bar{N}\rightarrow {\mbox{\rm End}}(V_\gamma)$ are analytic and the series converges in the spaces of smooth functions on $\bar{N}$ with values in ${\mbox{\rm End}}(V_\gamma)$. Inserting this expansion into (\ref{tyh}) we obtain $$(P\phi)(ka)=a^{-(\lambda+\rho)} \gamma(w) T (\hat{J}_\lambda \phi)(k) + \sum_{n\ge 1} a^{-(\lambda+\rho)-n\alpha_1} \psi_n(k)\ ,$$ where $$\psi_n(k):=\gamma(w)\int_{\bar{N}} A_n(\bar{n}) T a(\bar{n})^{\lambda-\rho }\phi(kw\kappa(\bar{n})) d\bar{n}\ .$$ Note that $k\mapsto \phi(kw\kappa(.))$ is a smooth family of distributions with compact support in $\bar{N}$. Thus $\psi_n$ is smooth. This finishes the proof of the lemma. \hspace*{\fill}$\Box$ \\[0.5cm]\noindent We now continue the proof of Lemma \ref{th43}. If ${\rm Re }(\lambda)<0$, then we reduce to the case ${\rm Re }(\lambda)>0$ replacing $\phi$ by $\hat{J}_\lambda(\phi)$. We can do this because $\lambda$ is then non-integral and $\hat{J}_\lambda$ is regular and bijective. Note that in this case $\chi_{\mu_\sigma+\rho_m-\lambda}$ is non-integral. Thus assume that ${\rm Re }(\lambda)\ge 0$. We choose the representation $\gamma\in \hat{K}$ and $T\in {\mbox{\rm Hom}}_M(V_\sigma,V_\gamma)$ such that $P:=P^T_\lambda$ is injective. The range of $P$ can be identified with the kernel of a certain invariant differential operator $D:C^{\infty}(X,V(\gamma))\rightarrow C^{\infty}(X,V(\gamma^\prime))$ for a suitable representation $\gamma^\prime$ of $K$ (see \cite{bunkeolbrich947}, Sec.3). In particular, $P\phi$ is real-analytic. We now assume $\phi\not\equiv 0$. Moreover, without loss of generality we can assume that $ M\in U$. Since $P\phi$ is real analytic, the expansion (\ref{epan}) does not vanish. Let $m$ be the smallest integer such that $\psi_m\not\equiv 0$ near $M$ (where $\psi_0:=\gamma(w)T\hat{J}_\lambda \phi)$. We argue that $m=0$ and thus obtain a contradiction. To prove that $m=0$ we again argue by contradiction. Assume that $m>0$. With respect to the coordinates $k,a$ the operator $D$ has the form $D=D_0+a^{-\alpha_1}R(a,k)$, where $D$ is a constant coefficient operator on $A$ and $R$ remains bounded if $a\to\infty$ (see \cite{warner721}, Thm. 9.1.2.4). Moreover, it is known that $D_0$ coincides with the $\bar{N}$-radial part of $D$. Choose $k\in K$ near $1$ and $\sigma^\prime\subset \gamma_{|M}$ such that that there exists an orthogonal projection $S\in {\mbox{\rm Hom}}_M(V_\gamma,V_{\sigma^\prime})$ with $S \gamma(k)\psi_m(k) =:v \not= 0$. Consider the $\bar{N}$-invariant section $f\in C^{\infty}(X,V(\gamma))$ defined by $$f(\bar{n}a):=a^{-(\lambda+\rho+m\alpha_1)} S^* v\ .$$ Since $D$ annihilates the asymptotic expansion (\ref{epan}), one can check that $Df=D_0f=0$ and thus $f=P\phi_1$ for some $\bar{N}$-invariant $\phi_1\in C^{-\infty}(\partial X,V(\sigma_\lambda))$. Now $f=P^S_{\lambda+m\alpha_1}\delta v$, where $\delta v\in C^{-\infty}(\partial X,V(\sigma^\prime_{\lambda+m\alpha_1}))$ is the delta distribution at $1$ with vector part $v$. Since $D$ and $P^S_{\lambda+m\alpha_1}$ are $G$-equivariant and $\delta v$ generates the $G$-module $C^{-\infty}(\partial X,V(\sigma^\prime_{\lambda+m\alpha}))$, we obtain a non-trivial intertwining operator $I$ from $C^{-\infty}(\partial X,V(\sigma^\prime_{\lambda+m\alpha}))$ to the kernel of $D$, hence to $C^{-\infty}(\partial X,V(\sigma_\lambda))$. Thus principal series representations $\pi^{\sigma,\lambda}$ and $\pi^{\sigma^\prime,\lambda+m\alpha_1}$ have the same infinitesimal character $\chi_{\mu_\sigma+\rho_m-\lambda}$. Since by assumption this character is non-integral both principal series are irreducible (see \cite{collingwood85}, 4.3.3). Hence $I$ is an isomorphism. Because of $m\not=0$ we have $\sigma\not=\sigma^\prime$. This implies that $\pi^{\sigma,\lambda}$ and $\pi^{\sigma^\prime,\lambda+m\alpha_1}$ can not be isomorphic. This is the contradiction we aimed at. We conclude that $m=0$.\hspace*{\fill}$\Box$ \\[0.5cm]\noindent The above argument can be extended to cover some cases of $\lambda\in{\bf a}^*$, ${\rm Re }(\lambda)\ge 0$, with $\chi_{\mu_\sigma+\rho_m-\lambda}$ integral. This would lead to stronger vanishing results for ${}^\Gamma C^{-\infty}(\Lambda,V(\sigma_\lambda))$ below. For example, if $\sigma$ is the trivial $M$-type, then the assertion of Lemma \ref{th43} holds true for all $\lambda\in {\aaaa_\C^\ast}$ with ${\rm Re }(\lambda)\ge 0$. But there exist examples of $\sigma\in\hat{M}$ and $\lambda\in{\bf a}^*$ with ${\rm Re }(\lambda)\ge 0$ and $\chi_{\mu_\sigma+\rho_m-\lambda}$ integral where the assertion of Lemma \ref{th43} is false. The possible failure of the lemma at $\lambda=0$ is connected with the existence of $L^2(Y,V_Y(\gamma))_{scat}$. \begin{kor}\label{ujn} Assume that $\lambda\in {\aaaa_\C^\ast}$, that ${\rm Re }(\lambda)\ge 0$, and that $\chi_{\mu_\sigma+\rho_m-\lambda}$ is non-integral. Then $$res_\Omega\circ \hat{J}_\lambda:C^{-\infty}(\Lambda,V(\sigma_\lambda))\rightarrow C^{\infty}(\Omega,V(\sigma_{-\lambda}))$$ is injective. \end{kor} \begin{lem}\label{ghu} Assume that $\lambda\in {\aaaa_\C^\ast}$, that ${\rm Re }(\lambda)\ge 0$, that $\chi_{\mu_\sigma+\rho_m-\lambda}$ is non-integral, and that $$ext:C^{\infty}(B,V_B(\sigma_\lambda))\rightarrow {}^\Gamma C^{-\infty}(\partial X,V(\sigma_\lambda))$$ is regular. Then ${}^\Gamma C^{-\infty}(\Lambda,V(\sigma_\lambda))=0$ . \end{lem} {\it Proof.$\:\:\:\:$} The assumption of the lemma implies that $$ext:C^{\infty}(B,V_B(\tilde{\sigma}_\lambda))\rightarrow {}^\Gamma C^{-\infty}(\partial X,V(\tilde{\sigma}_\lambda))$$ is regular. Indeed, if this extension is singular, then $\lambda$ is real and $$ext:C^{\infty}(B,V_B(\sigma_\lambda))\rightarrow {}^\Gamma C^{-\infty}(\partial X,V(\sigma_\lambda))$$ is singular since there is a conjugate linear isomorphism $\sigma\cong\tilde{\sigma}$. Let $\phi\in{}^\Gamma C^{-\infty}(\Lambda,V(\sigma_\lambda))$. Then for all $f\in C^{\infty}(B,V_B(\tilde{\sigma}_\lambda))$ we have by Lemma \ref{green} $$0=\langle res\circ \hat{J}_\lambda(\phi),res\circ ext(f)\rangle= \langle res\circ \hat{J}_\lambda(\phi), f \rangle \ .$$ Thus $res\circ \hat{J}_\lambda(\phi)=0$ and by Corollary \ref{ujn} we conclude $\phi=0$. \hspace*{\fill}$\Box$ \\[0.5cm]\noindent \begin{lem}\label{extregatim} If ${\rm Re }(\lambda)= 0$ and ${\rm Im}(\lambda)\not=0$, then $$ext:C^{-\infty}(B,V_B(\sigma_\lambda))\rightarrow {}^\Gamma C^{-\infty}(\partial X,V(\sigma_\lambda))$$ is regular. \end{lem} {\it Proof.$\:\:\:\:$} Note that $\chi_{\mu_\sigma+\rho_m-\lambda}$ is a non-integral character of ${\cal Z}$. Since $\lambda\in{\aaaa_\C^\ast}(\sigma)$, the extension is meromorphic at $\lambda$. Assume that $ext$ has a pole. The singular part of $ext$ maps to distributions which are supported on the limit set $\Lambda$. Then by Corollary \ref{ujn} the scattering matrix $S_\lambda=res\circ J_\lambda\circ ext$ has a pole at $\lambda$, too. But this contradicts Lemma \ref{unitary}. \hspace*{\fill}$\Box$ \\[0.5cm]\noindent In the following proposition we formulate the most complete vanishing result for\linebreak[4] ${}^\Gamma C^{-\infty}(\Lambda,V(\sigma_\lambda))$, ${\rm Re }(\lambda)\ge 0$, which is stated in the present paper. In general, it is not optimal for integral infinitesimal characters. We give upper bounds depending on $\delta_\Gamma$ for the parameters $\lambda$ with ${}^\Gamma C^{-\infty}(\Lambda,V(\sigma_\lambda))\not=0$. One can then deduce corresponding bounds for the discrete spectrum of ${\cal Z}$ on $L^2(Y,V_Y(\gamma))$. If one is only interested in the finiteness of the discrete spectrum, then it is sufficient to know the rough bound $\lambda \le \rho$, which is implied by the classification of the unitary representations of $G$ (see the proof of Corollary \ref{nahenull}). \begin{prop}\label{upperbound} Let ${\rm Re }(\lambda)\ge 0$. Then any of the following conditions implies that $${}^\Gamma C^{-\infty}(\Lambda,V(\sigma_\lambda))=0\ .$$ \begin{enumerate} \item ${\rm Im}(\lambda)\not=0$, \item ${\rm Re }(\lambda)>\delta_\Gamma$ and $\chi_{\mu_\sigma+\rho_m-\lambda}$ is a non-integral character of ${\cal Z}$, \item $$\hspace{-2cm}{\rm Re }(\lambda)>\frac{\rho^2+\rho\delta_\Gamma}{3\rho-\delta_\Gamma}$$ \end{enumerate} \end{prop} {\it Proof.$\:\:\:\:$} We consider $1$. If ${\rm Re }(\lambda)>0$, then the assertion follows from Lemma \ref{pokm}. If ${\rm Re }(\lambda)=0$, then we apply Lemmas \ref{ghu} and \ref{extregatim}. Sufficiency of condition $2.$ follows from Lemma \ref{ghu} and \ref{defofext}. Condition $3.$ is important for $\lambda$ with $\chi_{\mu_\sigma+\rho_m-\lambda}$ integral. In this case the relation of the space ${}^\Gamma C^{-\infty}(\Lambda,V(\sigma_\lambda))$ with the singularities of $ext$ is quite unclear. Let $\gamma$ denote a minimal $K$-type of the principal series representation $\pi^{\sigma,\lambda}$ of $G$ on $C^\infty(\partial X,V(\sigma_\lambda))$, and let $T\in {\mbox{\rm Hom}}_M(V_\sigma,V_\gamma)$ be an isometric embedding. Then the Poisson transform $$P^T_\lambda :C^{-\infty}(\partial X,V(\sigma_\lambda))\rightarrow C^\infty(X,V(\gamma))$$ is injective. Let $\psi\in {}^\Gamma C^{-\infty}(\Lambda,V(\sigma_\lambda))$ and $f:=P^T_\lambda\psi\in C^\infty(Y,V_Y(\gamma))$. Since ${\rm Re }(\lambda)>0$ we have for any test function $\phi\in C^\infty(\partial X,V(\sigma_{\bar{\lambda}}))$ $$\bar{c}_\sigma(\lambda)(\psi,\phi)=\lim_{a\to\infty} a^{\rho-\bar{\lambda}}\int_K (f(ka), T \phi(k)) dk \ .$$ We will show that condition $3.$ implies that $(\psi,\phi)=0$. Since $\phi$ was arbitrary, it follows $\psi=0$. In the following argument we assume that ${\rm Re }(\lambda)<\rho$. An easy modification works for ${\rm Re }(\lambda)\ge\rho$. The asymptotic expansion of $f$ near $\Omega$ obtained in Lemma \ref{eee} implies that $f\in L^p(Y,V_Y(\gamma))$ for $p=\frac{2\rho}{\rho+{\rm Re }(\lambda)}+\epsilon$, $\forall \epsilon>0$. If we set $\tilde{f}(ka)=a^{-(\rho+\delta_\Gamma+\epsilon)/p}f(ka)$, then $\tilde{f}\in L^p(X,V(\gamma))$. In fact we can estimate \begin{eqnarray*} \|\tilde{f}\|_{L^p(X,V(\gamma))}&=&\sum_{g\in\Gamma} \int_{gF} |\tilde{f}(g)|^p dg\\ &\le & \sum_{g\in \Gamma} \min\{a_h\:|\: h\in gF\}^{-(\rho+\delta_\Gamma+\epsilon)} \int_{gF} |f(g)|^p dg\\ &\le & C \sum_{g\in \Gamma} a_g^{-(\rho+\delta_\Gamma+\epsilon)}\\ &<& \infty \ . \end{eqnarray*}\ Let $\chi\in C^\infty_c(0,1)$ be such that $\int_0^1\chi(t) dt=1$. We extend $\chi$ to ${\bf R}$ by zero. Then we define $\chi_n\in C_c^\infty(A_+)$ by $\chi_n(a):=a^{-\rho-\bar{\lambda}}\chi(|\log(a)|-n)$. If we set $\phi_n(ka)= T\chi_n(a)\phi(k)$, then we can write $$\bar{c}_\sigma(\lambda)(\psi,\phi ) =\lim_{n\to \infty} (f,\phi_n) \ .$$ Now let $\tilde{\phi}_n(ka):= a^{(\rho+\delta_\Gamma+\epsilon)/p} \phi_n(ka)$. Then \begin{equation}\label{thishere} \bar{c}_\sigma(\lambda)(\psi,\phi) =\lim_{n\to \infty} (\tilde{f},\tilde{\phi}_n) \ .\end{equation} We consider $\Psi(ka):=a^{(\rho+\delta_\Gamma+\epsilon)/p}a^{-\rho-\bar{\lambda}}T \phi(k)$ and let $q\in (1,\infty)$ be the dual exponent to $p$. Note that $\bar{c}_\sigma(\lambda)\not=0$. If we assume that $\Psi\in L^q(X,V(\gamma))$, then $(\psi,\phi)=0$ follows from (\ref{thishere}). The following inequality implies that $\Psi\in L^q(X,V(\gamma))$: \begin{equation}\label{ineq}q\left(\frac{\rho+\delta_\Gamma+\epsilon}{p}-(\rho+{\rm Re }(\lambda))\right)<-2\rho \ .\end{equation} A simple computation shows that (\ref{ineq}) indeed follows from $3$. \hspace*{\fill}$\Box$ \\[0.5cm]\noindent If $\chi_{\mu_\sigma+\rho_m-\lambda}$ is integral, then one can combine the argument of Lemma \ref{th43}, or the understanding of its failure, respectively, with the above argument in order to choose a better exponent $p$. This leads to a stronger vanishing result. We omit this rather involved discussion, partly because we do not know how to formulate a general result. This omission causes some loss of information about the discrete spectrum of ${\cal Z}$ on $L^2(Y,V_Y(\gamma))$. In general, we do not know whether $ext$ is meromorphic at $\lambda=0$. But we can show the following result. \begin{prop}\label{mystic} If $\hat J_\lambda$ has a pole at $\lambda=0$, then $ext$ is meromorphic in a neighbourhood of $0$. \end{prop} The proof of the proposition will occupy the remainder of this section. First we want to fix an important corollary which applies to all $M$-types $\sigma$. \begin{kor}\label{nahenull} There exists $\epsilon>0$ such that $ext$ is regular and ${}^\Gamma C^{-\infty}(\Lambda,V(\sigma_\lambda))=0$ for $\lambda\in (0,\epsilon)$. \end{kor} {\it Proof.$\:\:\:\:$} If contrary to the assertion there is a sequence $\lambda_\alpha>0$ of poles of $ext$ with $\lambda_\alpha\to 0$, then the residues of $ext$ give elements of ${}^\Gamma C^{-\infty}(\Lambda,V(\sigma_{\lambda_\alpha}))$. Applying to these elements a suitable Poisson transform $P^T_{\lambda_\alpha}$, $T\in {\mbox{\rm Hom}}_M(V_\sigma,V_\gamma)$, we obtain smooth sections in $L^2(Y,V_Y(\gamma))$. Considered as elements of $L^2(\Gamma\backslash G)\otimes V_\gamma$ they generate unitary representations of $G$ whose underlying Harish-Chandra modules are dual to the underlying Harish-Chandra modules of $C^{\infty}(\partial X,V(\sigma_{\lambda_\alpha}))$, at least for $\lambda_\alpha$ sufficiently small. Here we have used that principal series representations with non-integral infinitesimal character are irreducible. Since the dual of a unitary representation is unitary too, we conclude that the principal series representations $C^{\infty}(\partial X,V(\sigma_{\lambda_\alpha}))$ carry invariant Hermitian scalar products or, in representation theoretic language, belong to the complementary series. But, by a result of Knapp-Stein \cite{knappstein71}, Par. 14, there is a complementary series for $\sigma$ iff $P_\sigma(0)=0$ or, equivalently, iff $\hat J_\lambda$ has a pole at $\lambda=0$. But in this case $ext$ is meromorphic at $\lambda=0$ by Proposition \ref{mystic}. This is in contradiction with the existence of the sequence $\lambda_\alpha$ of poles of $ext$ with $\lambda_\alpha\to 0$. \hspace*{\fill}$\Box$ \\[0.5cm]\noindent We now turn to the proof of Proposition \ref{mystic}. If $\delta_\Gamma<0$, there is nothing to show. For $\delta_\Gamma\ge 0$ we are going to use the embedding trick as in the proof of Proposition \ref{part1}. Let $G=G^n\subset G^{n+1}\subset\dots$ be the corresponding series of groups. Choose $k\in{\bf N}$ such that $\delta_\Gamma^{n+k}=\delta_\Gamma^{n}-\rho^{n+k}+\rho^n<0$. If there is a representation $\sigma^{n+k}$ of $M^{n+k}$ such that $\sigma^n:=\sigma\subset \sigma^{n+k}_{\ |M^n}$ and $-\rho^{n+k}+\rho^n \in {\aaaa_\C^\ast} (\sigma^{n+k})$, then $ext^{n+k}$ is meromorphic at $\lambda=-\rho^{n+k}+\rho^n$, which in turn implies the meromorphy of $ext^n$ at $\lambda=0$. Recall the definition (\ref{acagood}) of ${\aaaa_\C^\ast} (\sigma^{n+k})$. Since $\hat J_\lambda$ has at most first-order poles $$P_{\sigma^{n+k}}(-\rho^{n+k}+\rho^n)=0$$ implies that $-\rho^{n+k}+\rho^n \in {\aaaa_\C^\ast} (\sigma^{n+k})$. It follows from the functional equation (\ref{spex}) that the intertwining operator $\hat J_\lambda$ has a pole at $\lambda=0$ iff $P_{\sigma^n}(0)=0$. In this case $\sigma^n$ is irreducible and Weyl-invariant. So the discussion above shows that the following lemma implies Proposition \ref{mystic}. \begin{lem}\label{casebycase} Let $G^n$ one of the following four series of groups : \begin{eqnarray*} Spin(1,2n)&&n\ge 1\ ,\\ Spin(1,2n+1)&&n\ge 1\ ,\\ SU(1,n)&&n\ge 2\ ,\\ Sp(1,n)&&n\ge 2\ . \end{eqnarray*} If $\sigma^n \in \hat M^n$ (i.e. $\sigma^n$ is irreducible) is Weyl invariant and satisfies $P_{\sigma^n}(0)=0$, then for any $k\in{\bf N}$ there exists $\sigma^{n+k} \in \hat M^{n+k}$ with $\sigma^n\subset \sigma^{n+k}_{\ |M^n}$ and $P_{\sigma^{n+k}}(-\rho^{n+k}+\rho^n)=0$. \end{lem} {\it Proof.$\:\:\:\:$} First we need explicit expressions of the Plancherel densities $P_{\sigma^n}$ as given e.g. in \cite {knapp86}, Prop. 14.26. Consider the Cartan subalgebra ${\bf h}_{\bf C}^n={\bf a}_{\bf C}\oplus{\bf t}_{\bf C}^n$, ${\bf t}^n$ being a Cartan subalgebra of ${\bf m}^n$, and the subset $\Delta_r^n$ of the roots $\Delta^n$ of ${\bf h}_{\bf C}^n$ in ${\bf g}^n_{\bf C}$ given by $$ \Delta_r^n:=\{\alpha \in \Delta^n\:|\:\alpha_{|{\bf a}} \mbox{ is a root of ${\bf a}$ in }{\bf n}^n\}\ .$$ For $G^n\not=Spin(1,2n+1)$ there is exactly on real root $\beta_r\in \Delta_r^n$ distinguished by $\beta_{r\:|{\bf t}^n}=0$. Furthermore, in this case we consider a special element $m^n$ of the center of $M^n$. For $G=Spin(1,2n)$ the element $m^n$ is the non-trivial element in the kernel of the projection $Spin(1,2n)\rightarrow SO_0(1,2n)$. For $G^n=SU(1,n)$ $(Sp(1,n))$ we use the standard representation in $Gl(n+1,{\bf C})$ $(Gl(n+1,\bf H))$ in order to fix $m^n$ by $$ m^n:= \left( \begin{array}{rrc} -1&0&0\\ 0&-1&0\\ 0&0&{\mbox{\rm id}}_{n-1} \end{array} \right)\ . $$ Since $(m^n)^2=1$ we have $\sigma^n(m^n)=\pm {\mbox{\rm id}}$. The embedding $M^n\hookrightarrow M^{n+k}$ sends $m^n$ to $m^{n+k}$. There are nontrivial constants $C(\sigma^n)$ such that \begin{equation}\label{blanche} P_{\sigma^n}(\lambda)=C(\sigma^n)f_{\sigma^n}(\lambda)\prod_{\beta\in \Delta_r^n} \langle \mu_{\sigma^n}+\rho_m^n-\lambda,\beta\rangle\ , \end{equation} where $\langle.,.\rangle$ is a Weyl-invariant bilinear scalar product on $({\bf h}_{\bf C}^n)^*$ and $$ f_{\sigma^n}(\lambda)=\left\{ \begin{array}{ccl} 1&G^n=Spin(1,2n+1)&\\ \tan \left(\pi \frac{\langle\lambda,\beta_r\rangle}{\langle\beta_r,\beta_r\rangle}\right), &G^n\not=Spin(1,2n+1),& \sigma^n(m^n)=-{\rm e}^{2\pi\imath\frac{\langle\rho^n,\beta_r\rangle}{\langle\beta_r,\beta_r \rangle}}{\mbox{\rm id}}\\ \cot \left(\pi \frac{\langle\lambda,\beta_r\rangle}{\langle\beta_r,\beta_r\rangle}\right), &G^n\not=Spin(1,2n+1),& \sigma^n(m^n)={\rm e}^{2\pi\imath\frac{\langle\rho^n,\beta_r\rangle}{\langle\beta_r,\beta_r \rangle}}{\mbox{\rm id}} \end{array} \right. \ . $$ According to these three possibilities of the form of the Plancherel density we distinguish between the odd-dimensional, the tan and the cot case. We shall construct the representation $\sigma^{n+k}$ case by case. The easiest one is \noindent {\bf The tan case}\newline For any representation $\sigma^n$ fitting into this case we have $$P_{\sigma^n}(0)=\tan(0)=0\ .$$ Let $\sigma^{n+k}\in \hat M^{n+k}$ be an arbitrary representation satisfying $\sigma^n\subset \sigma^{n+k}_{\ |M^n}$. Then $\sigma^{n+k}(m^{n+k})=-{\mbox{\rm id}}$ iff $\sigma^n(m^n)=-{\mbox{\rm id}}$. Thus $\sigma^{n+k}$ belongs to the $\tan$-case iff $\rho^{n+k}-\rho^n$ is an integer multiple of $\beta_r$ (this is always the case except when $G^n=SU(1,n)$ and $k$ is odd). We conclude that $$ P_{\sigma^{n+k}}(-\rho^{n+k}+\rho^n)=\left\{ \begin{array}{c} C\tan(\pi\frac{\langle\-\rho^{n+k}+\rho^n,\beta_r\rangle}{\langle\beta_r,\beta_r \rangle})\\ C^\prime\cot(\pi\frac{\langle\-\rho^{n+k}+\rho^n,\beta_r\rangle}{\langle\beta_r,\beta_r \rangle}) \end{array}\right\} =0\ .$$ \noindent {\bf The odd-dimensional case}\newline As a Cartan subalgebra of ${\bf g}^n$ we choose $${\bf h}^n:= \left\{ T_\nu:= \left( {\scriptsize \begin{array}{cccc} \begin{array}{cc} 0&\nu_0\\ \nu_0&0 \end{array}&&&\\ &\begin{array}{cc} 0&-\nu_1\\ \nu_1&0 \end{array}&&\\ &&\ddots&\\ &&&\begin{array}{cc} 0&-\nu_n\\ \nu_n&0 \end{array} \end{array} } \right)\:\Bigg|\:\nu_i\in{\bf R} \right\}\ \ ,$$ where ${\bf a}=\{T_\nu\:|\:\nu_i=0,\:i=1,\dots,n\}$ and ${\bf t}^n=\{T_\nu\:|\:\nu_0=0\}$. Define $e_i\in ({\bf h}_{\bf C}^n)^*$ by $e_0(T_\nu):=\nu_0$ and $e_i(T_\nu):=\imath \nu_i$, $i=1,\dots,n$. We normalize $\langle.,.\rangle$ such that $\{e_i\}$ becomes an orthonormal basis of $({\bf h}_{\bf C}^n)^*$. Sometimes we will write elements of $\imath ({\bf t}^n)^*$ as $n$-tuples of reals with respect to the basis $e_1,\dots,e_n$. We choose positive roots of ${\bf t}^n$ in ${\bf m}^n_{\bf C}$ as follows: $$\Delta_m^{n,+}:=\{e_i\pm e_j \:|\: 1\le i<j\le n\}\ .$$ Then $\rho^n_m=(n-1,n-2,\dots,1,0)$ and the irreducible representations of $M^n$ correspond to the highest weights $$ \hat M^n \cong \{\mu_{\sigma^n}=(m_1,\dots,m_n)\:|\:m_1\ge\dots\ge m_{n-1}\ge |m_n|,\: m_i-m_j\in{\bf Z},\: m_n\in\frac{1}{2}{\bf Z}\}\ .$$ Furthermore $$\Delta^n_r=\{e_0\pm e_i\:|\:i=1,\dots,n\}\ .$$ We obtain $\rho^n=ne_0$ and for $\lambda=ze_0$ $$P_{\sigma^n}(\lambda)= C(\sigma^n)\prod_{i=1}^{n}(z^2-(m_i+n-i)^2)\ .$$ We see that $P_{\sigma^n}(0)=0$ iff $m_n=0$. In this case set $$\mu_{\sigma^{n+k}}:=(m_1,\dots,m_{n-1},m_n=0,\underbrace{0,\dots,0} _{\mbox{\scriptsize $k$ times}})\ .$$ Then $\sigma^n\subset \sigma^{n+k}_{\ |M^n}$ and the $n$-th factor of $P_{\sigma^{n+k}}$ is given by $z^2-(m_n+n+k-n)^2 = z^2- k^2$. Thus $P_{\sigma^{n+k}}(-\rho^{n+k}+\rho^n)=P_{\sigma^{n+k}}(-ke_0)=0$. \noindent {\bf The cot case}\\ First we observe that if $G=Spin(1,2n)$ and $\sigma^n$ belongs to the $\cot$-case, then $P_{\sigma^n}(0)\not=0$. In fact, in this case $\sigma^n$ is a faithful representation of $M^n=Spin(2n-1)$ and the only root $\beta\in\Delta^n_r$ perpendicular to $\mu_{\sigma^n}+\rho_m^n$ is $\beta_r$. We leave the simple verification to the reader. Since $\cot$ has a pole at $0$ the observation follows from (\ref{blanche}). An alternative proof is given in \cite{knappstein71}, Prop. 55. We are left with the discussion of $G^n=SU(1,n)$ and $G^n=Sp(1,n)$. We start with $G^n=SU(1,n)$. The group $M^n$ has the form $$M^n= \left\{\left(\begin{array}{ccc} z&0&0\\0&z&0\\0&0&B \end{array}\right)\:|\:\: z\in U(1),B\in U(n-1), z^2\det B=1\right\}\ .$$ We consider the Cartan subalgebra of ${\bf u}(1,n)$ $$\tilde{\bf h}^n:= \left\{ T_\nu:= \left( {\scriptsize \begin{array}{ccccc} \imath \nu_1&\nu_0&&&\\ \nu_0&\imath \nu_1&&&\\ &&\imath\nu_2&&\\ &&&\ddots&\\ &&&&\imath\nu_n \end{array} } \right)\:\Bigg|\:\nu_i\in{\bf R} \right\}\ \ ,$$ their subalgebras ${\bf a}:=\{T_\nu\:|\:\nu_i=0,\:i=1,\dots,n\}$ and $\tilde{\bf t}^n:=\{T_\nu\:|\:\nu_0=0\}$. Then ${\bf h}^n:=\{T_\nu\in\tilde{\bf h}^n\:|\:2\nu_1+\sum_{i=2}^n\nu_i=0\}$ and ${\bf t}^n:=\{T_\nu\in\tilde{\bf t}^n\:|\:2\nu_1+\sum_{i=2}^n\nu_i=0\}$ are Cartan subalgebras of ${\bf g}^n$ and ${\bf m}^n$, respectively. We define elements $\alpha,\beta,e_i\in (\tilde{\bf h}_{\bf C}^n)^*$, $i=2,\dots,n$, by $\alpha(T_\nu):=\nu_0$, $\beta(T_\nu)=\imath\nu_1$ and $e_i(T_\nu):=\imath \nu_i$. We extend $\langle.,.\rangle$ to $(\tilde{\bf h}^n_{\bf C})^*$ such that $\{\alpha,\beta,e_2,\dots,e_n\}$ becomes an orthogonal basis of $(\tilde{\bf h}_{\bf C}^n)^*$ with $|\alpha|^2=|\beta|^2=1$ and $|e_i|^2=2$. Sometimes we will write elements of $\imath (\tilde{\bf t}^n)^*$ as $n$-tuples of reals with respect to the basis $\beta,e_2,\dots,e_n$. We choose the positive roots of ${\bf t}^n$ in ${\bf m}^n_{\bf C}$ as follows: $$\Delta_m^{n,+}:=\{e_i-e_j \:|\: 2\le i<j\le n\}\ .$$ Then $\rho^n_m=\frac{1}{2}(0,n-2,n-4,\dots,4-n,2-n)$. Furthermore we have $$\Delta^n_r=\{2\alpha,\alpha\pm(\beta-e_i)\:|\:i=2,\dots,n\}$$ and $\rho^n=n\alpha$. We represent highest weights of representations of $M^n$ which are functionals on ${\bf t}^n$ by elements of $(\tilde{\bf t}^n_{\bf C})^*$: \begin{eqnarray*} \hat M^n &\cong& \{\mu_{\sigma^n}=(m_1,\dots,m_n)\:|\:m_2\ge\dots \ge m_n,\: m_i \in{\bf Z} \}\\ &&\quad\mbox{ modulo translation by elements of the form } (2\nu,\nu,\dots,\nu)\ . \end{eqnarray*} Then we can compute the scalar products with elements of $\Delta^n_r$ inside $(\tilde{\bf h}^n_{\bf C})^*$, and the result will not depend on the chosen representative. Since $\rho^n=n\alpha$ and $\beta_r=2\alpha$ we see that $\sigma^n$ belongs to the $\cot$-case iff $m_1\equiv n\:(2)$. We obtain for $\lambda=z\alpha$ $$P_{\sigma^n}(\lambda)= C(\sigma^n)\cot(\frac{\pi}{2}z)2z\prod_{i=2}^{n}(z^2-(m_1-n-2(m_i+1-i))^2)\ .$$ We see that $P_{\sigma^n}(0)=0$ iff for one index $i_0\in\{2,\dots,n\}$ the following equation holds: $$\frac{m_1-n}{2}=m_{i_0}+1-i_0\ .$$ In this case set $$\mu_{\sigma^{n+k}}:=(m_1,\dots,m_{i_0},\underbrace{m_{i_0},\dots,m_{i_0}} _{\mbox{\scriptsize $k$ times}},m_{i_0+1},\dots,m_n)\ .$$ Then $\sigma^n\subset \sigma^{n+k}_{\ |M^n}$. Depending on the parity of $k$ the representation $\sigma^{n+k}$ belongs to the $\tan$-case or $\cot$-case, respectively. In any case, the function $f_{\sigma^{n+k}}$ has a first-order pole at $\lambda=-\rho^{n+k}+\rho^n=-k\alpha$. But in addition to the $i_0$-th factor $z^2-(m_1-n-k-2(m_{i_0}+1-i_0))^2$ of the polynomial part of $P_{\sigma^{n+k}}(z\alpha)$ also the $(i_0+k)$-th factor $z^2-(m_1-n-k-2(m_{i_0}+1-i_0-k))^2$ is equal to $z^2- k^2$. Thus the polynomial part contributes a second order zero at $z=k$ and $P_{\sigma^{n+k}}(-\rho^{n+k}+\rho^n)=P_{\sigma^{n+k}}(-k\alpha)=0$. \begin{itemize}\item $G^n=Sp(1,n)$\end{itemize} The group $M^n$ has the form $$M^n= \left\{\left(\begin{array}{ccc} q&0&0\\0&q&0\\0&0&B \end{array}\right)\:|\:\: q\in Sp(1),B\in Sp(n-1)\right\}\ .$$ As a Cartan subalgebra of ${\bf g}^n$ we choose $${\bf h}^n:= \left\{ T_\nu:= \left( {\scriptsize \begin{array}{ccccc} \imath \nu_1&\nu_0&&&\\ \nu_0&\imath \nu_1&&&\\ &&\imath\nu_2&&\\ &&&\ddots&\\ &&&&\imath\nu_n \end{array} } \right)\:\Bigg|\:\nu_i\in{\bf R} \right\}\ \ ,$$ where ${\bf a}=\{T_\nu\:|\:\nu_i=0,\:i=1,\dots,n\}$ and ${\bf t}^n=\{T_\nu\:|\:\nu_0=0\}$. We define elements $\alpha,\beta,e_i\in ({\bf h}_{\bf C}^n)^*$, $i=2,\dots,n$, by $\alpha(T_\nu):=\nu_0$, $\beta(T_\nu)=\imath\nu_1$ and $e_i(T_\nu):=\imath \nu_i$. Then $\{\alpha,\beta,e_2,\dots,e_n\}$ becomes an orthogonal basis of $(\tilde{\bf h}_{\bf C}^n)^*$ and we normalize $\langle.,.\rangle$ such that $|\alpha|^2=|\beta|^2=1$ and $|e_i|^2=2$. Sometimes we will write elements of $\imath ({\bf t}^n)^*$ as $n$-tuples of reals with respect to the basis $\beta,e_2,\dots,e_n$. We choose the positive roots of ${\bf t}^n$ in ${\bf m}^n_{\bf C}$ as follows $$\Delta_m^{n,+}:=\{2\beta,e_i\pm e_j, 2e_i \:|\: 2\le i<j\le n\}\ .$$ Then $\rho^n_m=(1,n-1,n-2,\dots,2,1)$ and the irreducible representations of $M^n$ correspond to the highest weights $$ \hat M^n \cong \{\mu_{\sigma^n}=(m_1,\dots,m_n)\:|\:m_1\ge 0,m_2\ge\dots\ge m_{n}\ge 0,\: m_i\in{\bf Z}\}\ .$$ Furthermore we have $$\Delta^n_r=\{2\alpha,2(\alpha\pm\beta),\alpha\pm(\beta\pm e_i)\:|\:i=2,\dots,n\}$$ and $\rho^n=(2n+1)\alpha$. Since $\beta_r=2\alpha$ we see that $\sigma^n$ goes with cot iff $m_1$ is odd. We obtain for $\lambda=z\alpha$ \begin{eqnarray*} P_{\sigma^n}(\lambda)&=& C(\sigma^n)\cot(\frac{\pi}{2}z)\:2z\:4(z^2-(m_1+1)^2)\\ &&\quad \prod_{i=2}^{n}(z^2-(m_1+1+2(m_i+n+1-i))^2)(z^2-(m_1+1-2(m_i+n+1-i))^2)\ . \end{eqnarray*} We see that $P_{\sigma^n}(0)=0$ iff for one index $i_0\in\{2,\dots,n\}$ the following equation holds: $$\frac{m_1+1}{2}=m_{i_0}+n+1-i_0\ .$$ We are going to define $\sigma^{n+k}$ by an inductive procedure. It rests on the following claim. Let $l\in{\bf N}_0$ and $\mu_{\sigma^{n+l}}=(m^\prime_1,\dots,m^\prime_{n+l})$ be a highest weight of an irreducible representation of $M^{n+l}$ such that $m_1^\prime=m_1$, $m_{i_0}^\prime=m_{i_0}$ and one of the following conditions holds: \begin{enumerate} \item There exists an index $j_l\in\{i_0+l,\dots,n+l\}$ such that $$ m_1+1-2(m^\prime_{j_l}+n+l+1-j_l)=2l\ .$$ \item $m_1+1=2l$ . \item There exists an index $j_l\in\{i_0,\dots,n+l\}$ such that $$ m_1+1+2(m^\prime_{j_l}+n+l+1-j_l)=2l\ .$$ \end{enumerate} Then there exists $\sigma_{n+l+1}\in\hat M^{n+l+1}$ with the same properties ($l$ replaced by $l+1$) such that $\sigma^{n+l}\subset\sigma^{n+l+1}_{\ |M^{n+l}}$. We now prove the claim. If $\sigma^{n+l}$ satisfies condition 1 and $m^\prime_{j_l}> m^\prime_{j_l+1}$ (by convention $m^\prime_j:=0$ for $j>n+l$) we set $$\mu_{\sigma^{n+l+1}}:=(m^\prime_1,\dots,m^\prime_{j_l},m^\prime_{j_l}-1, m^\prime_{j_l+1},\dots,m^\prime_{n+l})\:,\ j_{l+1}:=j_l+1\ .$$ Then $\sigma^{n+l+1}$ also satisfies condition 1. If $\sigma^{n+l}$ satisfies condition 1 and $m^\prime_{j_l}= m^\prime_{j_l+1}$ we set $$\mu_{\sigma^{n+l+1}}:=(m^\prime_1,\dots,m^\prime_{j_l},m^\prime_{j_l}, m^\prime_{j_l}, m^\prime_{j_l+2},\dots,m^\prime_{n+l})\ .$$ If in addition $j_l<n+l$, set $j_{l+1}:=j_l+2$. Then again $\sigma^{n+l+1}$ satisfies condition 1. Otherwise we have $m^\prime_{n+l}=0$, hence $m_1-1=2l$. It follows that $\sigma^{n+l+1}$ satisfies condition 2. If $\sigma^{n+l}$ satisfies condition 2 or 3, then $\sigma^{n+l+1}$ defined by $$\mu_{\sigma^{n+l+1}}:=(m^\prime_1,\dots,m^\prime_{n+l},0)$$ satisfies condition 3 with $j_{l+1}=n+l+1$ or $j_{l+1}=j_l$, respectively. The branching rules for the restriction from $Sp(n+l+1)$ to $Sp(n+l)$ (see e.g. \cite{zelobenko73}, Ch. XVIII) show that in any case $\sigma^{n+l}\subset \sigma^{n+l+1}_{\ |M^{n+l}}$. The claim now follows. Since $\sigma^n$ satisfies the induction hypothesis of the claim for $l=0$ and $j_0=i_0$ we can define $\sigma^{n+k}$ inductively. We have to check that $P_{\sigma^{n+k}}(-\rho^{n+k}+\rho^n)=P_{\sigma^{n+k}}(-2k\alpha)=0$. In fact, the claim ensures that $P_{\sigma^{n+k}}(z\alpha)$ containes in addition to $z^2-(m_1+1-2(m_{i_0}+n+k+1-i_0))^2$ a second factor which contributes with $z^2-(2k)^2$. Thus the pole originating from the $\cot$ factor cancels, and we have $P_{\sigma^{n+k}}(-2k\alpha)=0$. Now the lemma and, hence, Proposition \ref{mystic} is proved. \hspace*{\fill}$\Box$ \\[0.5cm]\noindent \section{The essential spectrum}\label{esess} In the present section we consider the spectral comparison between $L^2(X,V(\gamma))$ and $L^2(Y,V_Y(\gamma))$. Let ${\cal A}$ be any commutative algebra of invariant differential operators on $V(\gamma)$ which is generated by selfadjoint elements and contains ${\cal Z}_\gamma$. As for ${\cal Z}_\gamma$ there are spectral decompositions of $L^2(X,V(\gamma))$ and $L^2(Y,V_Y(\gamma))$ with respect to ${\cal A}$. The main result is that the essential spectrum of ${\cal A}$ on $L^2(X,V(\gamma))$ and $L^2(Y,V_Y(\gamma))$ coincides. Recall the characterization of the essential spectrum in terms of Weyl sequences. Let $\{A_i\}$ be a finite set of generators of ${\cal A}$. A character $\lambda$ of ${\cal A}$ belongs to the essential spectrum of ${\cal A}$ iff there exists a Weyl sequence $\{\phi_\alpha\}\subset C^\infty_c$ (i.e. a sequence without accumulation points in $L^2$) such that $\max_i \|A_i\phi_\alpha-\lambda(A_i)\phi_\alpha\|_{L^2}\rightarrow 0$ when $\alpha\to \infty$. \begin{prop}\label{esspec} The essential spectrum of ${\cal A}$ on $L^2(X,V(\gamma))$ and on $L^2(Y,V_Y(\gamma))$ coincides. \end{prop} {\it Proof.$\:\:\:\:$} The proof of the proposition relies on the transfer of Weyl sequences. Let $\lambda$ be in the essential spectrum of ${\cal A}$ on $L^2(Y,V_Y(\gamma))$. Then there is a Weyl sequence $\{\phi_\alpha\}$, satisfying \begin{itemize} \item $\|\phi_\alpha\|_{L^2(Y,V_Y(\gamma))}=1$, $\forall \alpha$ \item $\{\phi_\alpha\}$ has no accumulation point in $L^2(Y,V_Y(\gamma))$, and \item $\max_i \| A_i\phi_\alpha-\lambda(A_i)\phi_\alpha\|_{L^2(Y,V_Y(\gamma))}\rightarrow 0$ as $\alpha\to \infty$. \end{itemize} Using a construction of Eichhorn \cite{eichhorn91} we can modify this Weyl sequence such that it satisfies in addition $\|\phi_\alpha\|_{L^2(K,V_Y(\gamma))}\rightarrow 0$ as $\alpha\to \infty$ for any compact $K\subset Y$. Let $\{V_i\}$ be a finite number of open subsets covering $Y$ at infinity such that each $V_i$ has a diffeomorphic lift $\tilde{V}_i\subset X$. Using the method of Lemma \ref{lll} we can choose the $V_i$ such that there exists a subordinated partition of unity $\{\chi_i\}$ (at infinity of $Y$) such that for $A\in {\cal A}$ $$|[A,\chi_i](ka)| \le C(A) a^{-1},\quad \forall kaK\in \tilde{V}_i\ .$$ By taking a subsequence of the Weyl sequence and renumbering the $V_i$ we can assume that $\|\chi_1 \phi_\alpha\|_{L^2(Y,V_Y(\gamma))}\ge c$ for some $c>0$ independent of $\alpha$. We set $\psi_\alpha:=\chi_1\phi_\alpha /\|\chi_1 \phi_\alpha\|_{L^2(Y,V_Y(\gamma))}$. We claim that $\psi_\alpha$ is again a Weyl sequence for $\lambda$. In fact $\|\psi_\alpha\|_{L^2(Y,V_Y(\gamma))}=1$ by definition, $\| \psi_\alpha\|_{L^2(K,V_Y(\gamma))}\rightarrow 0$ as $\alpha\to \infty$ for any compact $K\subset Y$. This implies that $\{\psi_\alpha\}$ has no accumulation points. It remains to verify that for $A\in {\cal A}$ $$\|(A-\lambda(A))\psi_\alpha\|_{L^2(Y)}\rightarrow 0, \quad \alpha\to \infty\ .$$ We have $$(A-\lambda(A))\psi_\alpha=\frac{\chi_1}{\|\chi_1 \phi_\alpha\|_{L^2(Y,V_Y(\gamma))}}(A-\lambda(A))\phi_\alpha +\frac{[A,\chi_1]\phi_\alpha}{\|\chi_1 \phi_\alpha\|_{L^2(Y,V_Y(\gamma))}} \ .$$ Obviously we have $$\|\frac{\chi_1}{\|\chi_1 \phi_\alpha\|_{L^2(Y,V_Y(\gamma))}}(A-\lambda(A))\phi_\alpha\|_{L^2(Y,V_Y(\gamma))}\rightarrow 0, \quad \alpha\to \infty\ .$$ For any $\epsilon>0$ we can choose $K\subset Y$ compact such that $\sup_{x\in Y\setminus K} |[A,\chi_1](x)|<\epsilon$. We then have \begin{eqnarray*} \lefteqn{\lim_{\alpha\to\infty}\|\frac{1}{\|\chi_1 \phi_\alpha\|_{L^2(Y,V_Y(\gamma))}}[A,\chi_1]\phi_\alpha\|_{L^2(Y,V_Y(\gamma))}}\hspace{1cm}\\ &\le &\lim_{\alpha\to\infty} \left( \frac{\sup_{x\in K}|[A,\chi_1](x)|}{\|\chi_1 \phi_\alpha\|_{L^2(Y,V_Y(\gamma))}} \|\phi_\alpha\|_{L^2(K,V_Y(\gamma))} + \frac{\epsilon}{\|\chi_1 \phi_\alpha\|_{L^2(Y,V_Y(\gamma))}} \|\phi_\alpha\|_{L^2(Y\setminus K,V_Y(\gamma))}\right) \\ &\le& \epsilon/c\ . \end{eqnarray*} It follows that $$\lim_{\alpha\to\infty}\|\frac{[A,\chi_1]\phi_\alpha}{\|\chi_1 \phi_\alpha\|_{L^2(Y,V_Y(\gamma))}} \|_{L^2(Y,V_Y(\gamma))}=0\ .$$ Thus $\{\psi_\alpha\}$ is a Weyl sequence of the algebra ${\cal A}$ with respect to the character $\lambda$ which is supported in $V_1$. Lifting this sequence to $X$ we obtain a Weyl sequence of ${\cal A}$ to $\lambda$ in $L^2(X,V(\gamma))$. Thus the essential spectrum of ${\cal A}$ on $L^2(Y,V_Y(\gamma))$ is contained in the essential spectrum of ${\cal A}$ on $L^2(X,V(\gamma))$. In order to prove the opposite inclusion we choose a finite cover of infinity of $X$ by sets $W_j$ such that the central projection of $W_j$ is not surjective onto $\partial X$. If the character $\lambda$ is in the essential spectrum of ${\cal A}$ on $L^2(X,V(\gamma))$, then repeating the above construction we can find a Weyl sequence $\{\psi_\alpha\}$ of ${\cal A}$ to $\lambda$ with ${\mbox{\rm supp}}(\psi_\alpha)\subset W_1$. There is a $g\in G$ such that $gW_1\subset V_1$. Then $g^*\psi_\alpha$ can be pushed down to $Y$ and gives a Weyl sequence for ${\cal A}$ to $\lambda$ on $L^2(Y,V_Y(\gamma))$. Thus the essential spectrum of ${\cal A}$ on $L^2(X,V(\gamma))$ is contained in the essential spectrum of ${\cal A}$ on $ L^2(Y,V_Y(\gamma))$. \hspace*{\fill}$\Box$ \\[0.5cm]\noindent \section{Relevant generalized eigenfunctions}\label{relsec} In principle a spectral decomposition of $L^2(Y,V_Y(\gamma))$ with respect to ${\cal Z}_\gamma$ is a way of expressing elements of $L^2(Y,V_Y(\gamma))$ in terms of generalized eigensection of ${\cal Z}$, i.e., sections in $C^\infty(Y,V_Y(\gamma))$ on which ${\cal Z}$ acts by a character. Its turns out that only a small portion of these eigenfunctions is are needed for the spectral decomposition. We call them relevant. Relevant eigenfunctions satisfy certain growth conditions. In order to deal with these growth conditions properly we introduce the Schwartz space $S(Y,V_Y(\gamma))$. This Schwartz space is the immediate generalization of Harish-Chandra's Schwartz space $S(X,V(\gamma))$ to our locally symmetric situation. The significance of the Schwartz space is the following. If a generalized eigenfunction is relevant for the spectral decomposition of $L^2(Y,V_Y(\gamma))$ with respect to ${\cal Z}_\gamma$ then it defines a continuous functional on the Schwartz space. The Schwartz space is defined as a Fr\'echet space which is contained in $L^2(Y,V_Y(\gamma))$. Then its (hermitian) dual $S(Y,V_Y(\gamma))^\prime$ contains the relevant generalized eigenfunctions. We now turn to the definition of $S(Y,V_Y(\gamma))$. The idea is to require the similar conditions as for $S(X,V(\gamma))$ locally at infinity. Let $W\subset X\cup\Omega$ be any compact subset. For any $A\in{\cal U}({\bf g})$, $N\in {\bf N}$, we define the seminorm $q_{W,A,N}(f)\in[0,\infty]$ of $f\in C^\infty(X,V(\gamma))$ by \begin{equation}\label{swart}q_{W,A,N}(f):=\int_{(W\cap X)K} \log (\|g\|^N ) |f(Ag)|^2 dg\ .\end{equation} Here $\|g\|$ denotes the norm of ${\mbox{\rm Ad}}(g)$ on ${\bf g}$. In the following definition we identifiy $C^\infty(Y,V_Y(\gamma))$ with the subspace of $\Gamma$-invariant sections in $C^\infty(X,V(\gamma))$. \begin{ddd} The Schwartz space is the space of sections $f\in C^\infty(Y,V_Y(\gamma))$ with $q_{W,A,N}(f)<\infty$ for all $W$, $A$ and $N$ as above. The seminorms $q_{W,A,N}$ define the Fr\'echet topology of $S(Y,V_Y(\gamma))$. \end{ddd} \begin{ddd} An eigenvector of ${\cal Z}_\gamma$ belonging to the dual of the Schwartz space $S(Y,V_Y(\gamma))^\prime$ is called tempered. \end{ddd} The main goal of the present section is to provide a list of all tempered eigenvectors of ${\cal Z}$. Unfortunately there exist exceptional characters where such a description is difficult. Fortunately this exceptional set is at most countable. We will cover the set of all characters of ${\cal Z}$ which may provide difficulties by a countable set $PS$ below. The set of $\lambda\in{\bf h}_{\bf C}^*$ of parametrizing integral characters of ${\cal Z}$ forms a lattice and is hence countable. We denote that set of integral characters of ${\cal Z}$ by $PS_i$. We choose a Cartan algebra ${\bf t}$ of ${\bf m}$ such that ${\bf a}\oplus{\bf t}={\bf h}$ and a positive root system of ${\bf t}$. Let $\rho_m$ denote half of the sum of the positive roots of $({\bf m},{\bf t})$. For $\sigma\in \hat{M}$ let $\mu_\sigma\in{\bf t}^*$ be the highest weight. Then the character $\chi_\lambda$ of ${\cal Z}$ on the principal series representation $C^\infty(\partial, V(\sigma_\mu))$ is parametrized by $\lambda:=\mu_\sigma+\rho_m-\mu\in{\bf h}_{\bf C}^*$. We observe that if $\chi_\lambda\not\in PS_i$, then $\mu\in{\aca(\sigma)}$. Indeed, a pole of $P_\sigma$ at $\mu$ implies the integrality $\chi_\lambda$ (compare (\ref{blanche})), whereas a pole of $\hat J_{-\mu}$ for non-integral $\chi_\lambda$ cancels with a zero of $P_\sigma$ at $\mu$ because of (\ref{spex}) and the irreducibility of $C^{\infty}(\partial X,V(\sigma_{\mu}))$. If $\chi\not\in PS_i$, then we have a complete understanding of the corresponding eigenspace $${\cal E}_\chi:= \{f\in C^\infty_{mg}(X,V(\gamma)), \quad (A-\chi(A))f=0 \quad \forall A\in{\cal Z} \}\ .$$ Let $(V_\gamma)_{|M}=\oplus_i V_\gamma(\sigma_i)$ denote the decomposition of $(V_\gamma)_{|M}$ into isotypic components. Set $r_i=[(V_\gamma)_{|M}:V_{\sigma_i}]$ and choose isomorphisms $T_i\in {\mbox{\rm Hom}}_M(\oplus_{j=1}^{r_i} V_{\sigma_i},V_\gamma(\sigma_i))$. Employing the fact that principal series representations with non-integral infinitesimal character are irreducible it is a simple matter to deduce the following lemma from the results of \cite{olbrichdiss}. \begin{lem}\label{oppp} If $\chi\not\in PS_i$, then the eigenspace ${\cal E}_\chi$ is the isomorphic image of the Poisson transform $$\bigoplus_i P_{\mu_i}^{T_i}:\bigoplus_i \bigoplus_{j=1}^{r_i}C^{-\infty}(\partial X,V(\sigma_{i,\mu_i}))\rightarrow C^\infty_{mg}(X,V(\gamma))\ ,$$ where $\mu_i\in{\aaaa_\C^\ast}$ is uniquely characterized by $\chi_{\mu_{\sigma_i}+\rho_m-\mu_i}=\chi$, ${\rm Re }(\mu_i)>0$ or ${\rm Re }(\mu_i)=0$ and ${\rm Im}(\mu_i)\ge 0$. \end{lem} Let $PS_d$ denote the set of all characters $\chi$ of ${\cal Z}$ such that there exists $\sigma\subset\gamma_{|M}$ and $\mu\in{\aaaa_\C^\ast}$ with ${\rm Re }(\mu)\ge 0$, $\chi_{\mu_\sigma+\rho_m-\mu}=\chi$ and ${}^\Gamma C^{-\infty}(\Lambda,V(\sigma_\mu))\not= 0$ or $\mu =0$. By Lemma \ref{ghu} and the fact that $ext$ is meromorphic on ${\aca(\sigma)}$ the set $PS_d$ is countable. Let $PS=PS_i\cup PS_d$. Then $PS$ is countable, too. We recall the asymptotic expansion of eigenfunctions (see \cite{wallach88}, Ch. 4, \cite{wallach92}, Ch.11). Since we want to avoid hyperfunction boundary values we consider eigensections of moderate growth. A section $f\in C^\infty(X,V(\gamma))$ is of moderate growth if for any $A\in {\cal U}({\bf g})$ there exists $R\in{\bf R}$ such that $$\sup_{g\in G} \|g\|^{-R}|f(Ag)| <\infty\ .$$ Let $C_{mg}^\infty(X,V(\gamma))$ be the space of all sections of $V(\gamma)$ of moderate growth. Let $L^+\subset {\bf a}^*_+$ be the semigroup generated by the positive roots of $({\bf a},{\bf n})$ and $0$. Let $f\in C^\infty(X,V(\gamma))$ be some ${\cal Z}$-finite section which is $K$-finite with respect to the action of $K$ by left translations. Then there is a finite set of leading exponents $E(f)\subset {\aaaa_\C^\ast}$ such that $$f(ka)\stackrel{a\to\infty}{\sim} \sum_{\mu\in E(f)} a^{\mu-\rho} \sum_{{\bf Q}\in L^+} a^{-Q} p(f,\mu,Q)(k)(\log(a))\ ,$$ where $p(f,\mu,Q)$ is a polynomial (required to be non-trivial for $Q=0$) on ${\bf a}$ with values in $C^\infty(K,V_\gamma)$. To read this expansion properly consider $f$ as a function on $G$ with values in $V_\gamma$. The leading coefficient of the polynomial $p(f,\mu,0)$ has a continuous extension with respect to $f$ which is a $G$-equivariant continuous map from the closed $G$-submodule of $C^\infty_{mg}(X,V(\gamma))$ generated by $f$ to $C^{-\infty}(\partial X,V(\gamma_{|M,\mu}))$. If $f$ is an eigensection of ${\cal Z}$ and $\mu\not=0$, then $p(f,\mu,0)$ is in fact a constant polynomial \cite{olbrichdiss}, Lemma 4.6. If $f\in C^\infty(Y,V_Y(\gamma))$ is a tempered generalized eigensection of ${\cal Z}$, then its $\Gamma$-invariance implies that $f\in C_{mg}^\infty(X,V(\gamma))$. It follows that $p(f,\mu,0)\in {}^\Gamma C^{-\infty}(\partial X,V(\gamma_{|M,\mu}))$ is a $\Gamma$-invariant distribution. \begin{lem}\label{wo} If $f$ is a tempered generalized eigensection of ${\cal Z}$, then ${\mbox{\rm supp}} (p(f,\mu,0))\subset\Lambda$ for all $\mu\in E(f)$ with ${\rm Re }(\mu)>0$. \end{lem} {\it Proof.$\:\:\:\:$} We argue by contradiction. Consider the exponent $\mu\in E(f)$ with the largest real part ${\rm Re }(\mu)>0$ such that ${\mbox{\rm supp}}(p(f,\mu,0))\cap\Omega \not= \emptyset$. We assume that such an exponent $\mu$ exists. Note that $p(f,\mu,0)$ is a constant polynomial on ${\bf a}$. We study the support of of $p(f,\mu,0)$ by testing this distribution against suitable test functions. Let $F\subset X\cup\Omega$ be a fundamental domain of $\Gamma$ and $\partial F:=F\cap\Omega$. Since $p(f,\mu,0)$ is $\Gamma$-invariant and since we have freedom to choose $F$ it suffices to show that ${\mbox{\rm supp}}(p(f,\mu,0))\cap{\rm int}(\partial F)=\emptyset$. Thus let $\phi\in C_c^\infty({\rm int}(\partial F),V(\gamma_{|M,-\bar{\mu}}))$ be a test function. The application of the distribution boundary value $p(f,\mu,0)$ to $\phi$ can be written as the limit $$ (\phi,p(f,\mu,0)) = \lim_{a\to \infty} a^{\rho-\mu} \int_K ( \phi(k), f(ka) )$$ (we write sesquilinear pairings as $(.,.)$). Let $\chi\in C^\infty_c(0,1)$ be such that $\int_0^1\chi(t) dt=1$. We extend $\chi$ to ${\bf R}$ by zero. Then we define $\chi_n\in C_c^\infty(A_+)$ by $\chi_n(a):=a^{-\rho-\bar{\mu}}\chi(|\log(a)|-n)$. If we set $\phi_n(ka)= \chi_n(a)\phi(k)$, then we can write \begin{equation}\label{mnb}( \phi,p(f,\mu,0)) =\lim_{n\to \infty} (\phi_n,f) \ .\end{equation} If $n$ is sufficiently large, then ${\mbox{\rm supp}}(\phi_n)\subset F$. $\phi_n$ descends to a Schwartz space section $\tilde{\phi}_n\in S(Y,V_Y(\gamma))$. Using the fact that ${\rm Re }(\mu)>0$ one can easily check that $\lim_{n\to \infty}\tilde{\phi}_n=0$ in $S(Y,V_Y(\gamma))$. The right-hand side of (\ref{mnb}) can be written as the application of $f\in S(Y,V_Y(\gamma))^\prime$ to $\tilde{\phi}_n\in S(Y,V_Y(\gamma))$. It follows that $$( \phi,p(f,\mu,0))=\lim_{n\to 0}(\tilde{\phi}_n,f)=0\ .$$ Since $\phi$ was arbitrary we conclude that ${\mbox{\rm supp}}(p(f,\mu,0))\cap{\rm int}(\partial F)=\emptyset$. As noted above it follows that ${\mbox{\rm supp}}(p(f,\mu,0))\cap\Omega=\emptyset$ and this contradicts our assumption.\hspace*{\fill}$\Box$ \\[0.5cm]\noindent The following proposition is a part of our description of the generalized eigenfunctions of ${\cal Z}$ which are relevant for the spectral decomposition. We adopt the notation of Lemma \ref{oppp}. \begin{prop}\label{gener1} If $\chi\not\in PS$ and $f\in {\cal E}_\chi\cap S(Y,V_Y(\gamma))^\prime$ is a tempered eigenfunction of ${\cal Z}$, then $f=\bigoplus_i P_{\mu_i}^{T_i}(\phi_i)$, where ${\rm Re }(\mu_i)\ge 0$, $\chi_{\mu_{\sigma_i}+\rho_m-\mu_i}=\chi$. Moreover, $\phi_i\not= 0$ implies that ${\rm Re }(\mu_i)=0$. \end{prop} {\it Proof.$\:\:\:\:$} Assume that $\chi\not\in PS$. Let $0\not=f\in E_\chi$ be a tempered eigenfunction of ${\cal Z}$. Then by Lemma \ref{oppp} we can represent $f$ as $\sum_i P_{\mu_i}^{T_i}(\phi_i)$. Here $\phi_i\in \bigoplus_{j=1}^{r_i}{}^\Gamma C^{-\infty}(\partial X,V(\sigma_{i,\mu_i}))$ is uniquely characterized by $c_\gamma(\mu_i)T_i\phi_i=p(f,\mu_i,0)$. If $\phi_i\not=0$, then by Lemma \ref{wo} and the definition of $PS_d\subset PS$ we have ${\rm Re }(\mu_i)=0$.\hspace*{\fill}$\Box$ \\[0.5cm]\noindent \section{Wave packets and scalar products}\label{wxa} In this section we introduce wave packets of Eisenstein series and show that they belong to the Schwartz space. Thus the scalar product of a wave packet with such a tempered generalized eigenfunction of ${\cal Z}$ makes sense. If the corresponding character does not belong to the exceptional set $PS$, then we obtain an explicit formula for this scalar product. The subspace $L^2(Y,V_Y(\gamma))_c\subset L^2(Y,V_Y(\gamma))$ spanned by the wave packets is the absolute contiuous subspace. We show that its complement is the discrete subspace $L^2(Y,V_Y(\gamma))_d$ and that there is no singular continuous subspace. It turns out that the support of the continuous spectrum of ${\cal Z}$ on $L^2(Y,V_Y(\gamma))$ concides with the support of the continuous spectrum of ${\cal Z}$ on $L^2(X,V(\gamma))$ which is well known by the Harish-Chandra Plancherel theorem. The main result of the present section is Theorem \ref{contsp}. First we introduce the notion of an Eisenstein series. Let $\sigma\in \hat{M}$, $T\in{\mbox{\rm Hom}}_M(V_\sigma,V_\gamma)$, and let $P^T_\mu$, $\mu\in{\aaaa_\C^\ast}$, be the associated Poisson transform (see Definition \ref{defofpoi}). \begin{ddd} The Eisenstein series associates to $\phi\in C^{-\infty}(B,V(\sigma_\mu))$ the eigensection of ${\cal Z}$ $$E(\mu,\phi,T):= P_\mu^T\circ ext(\phi)\in C^\infty(Y,V_Y(\gamma))$$ corresponding the character $\chi_{\mu_\sigma+\rho_m-\mu}$. \end{ddd} The meromorphic continuation of $ext$ immediately implies the following corollary. \begin{kor} For $\mu\in{\aca(\sigma)}$ the Eisenstein series $E(\mu,.,T):C^{-\infty}(B,V(\sigma_\mu))\rightarrow C^\infty(Y,V_Y(\gamma))$ is a meromorphic family of operators with finite-dimensional singularities. The Eisenstein series is holomorphic on $\{{\rm Re }(\mu)=0,\mu\not=0\}$. Moreover, $E(\mu,.,T)$ has at most first-order poles in the set $\{\lambda\in{\aca(\sigma)}\:|\:{\rm Re }(\mu)>0\}$. \end{kor} We now discuss the functional equations satisfied by the Eisenstein series. This functional equation will be deduced from the corresponding functional equation of the Poisson transform. Recall the definition (\ref{cgamma}) of $c_\gamma(\lambda)$. Next we recall the functional equation satisfied by the Poisson transform which is proved in \cite{olbrichdiss}: \begin{equation}\label{mi9}c_{\sigma}(\lambda) P^T_\lambda \circ J_{-\lambda} = P^{\gamma(w)c_{ \gamma}(\lambda)T}_{-\lambda } \ ,\end{equation} where $w\in N_K(M)$ represents the generator of the Weyl group of $({\bf g},{\bf a})$. If we use $J_\lambda\circ ext = ext\circ S_\lambda$, then we obtain the following corollary. \begin{kor}\label{funeq} The Eisenstein series satisfies the functional equation $$E(\lambda,c_{\sigma}(\lambda)S_{-\lambda}\phi,T)=E(-\lambda ,\phi,\gamma(w)c_{ \gamma}(\lambda)T)\ .$$ (To be more precise, this is an identity of meromorphic quantities valid for all $\lambda\in {\aaaa_\C^\ast}$ where all terms are meromorphic.) \end{kor} Now we turn to the definition of the wave packet transform. Roughly speaking, a wave packet of Eisenstein series is an average of the Eisenstein series over imaginary parameters with a smooth, compactly supported weight function. More precisely, the space of such weight functions ${\cal H}_0$ is the linear space of smooth families ${\bf a}^*_+\ni \mu\mapsto \phi_{\imath\mu}\in C^\infty(B,V_B(\sigma_{\imath\mu}))$ with compact support with respect to $\mu$. Because of the symmetry \ref{funeq} it will be sufficient to consider wave packets on the positive imaginary axis, only. Next we fix a convenient choice of the endomorphisms $T$ entering the definition of the Eisenstein series. Let $\tilde{\gamma}$ be the dual representation of $\gamma$ and let $\tilde{T}\in {\mbox{\rm Hom}}_M(V_{\tilde{\sigma}},V_{\tilde{\gamma}})$ be such that $\tilde{T}^*T={\mbox{\rm id}}$. We set $T(\lambda):=c_\sigma(\lambda)^{-1}T$ and $\tilde{T}(\lambda):=c_{\tilde{\sigma}}(\lambda)^{-1} \tilde{T}$. There is a conjugate linear isomorphism of $\sigma$ and $\tilde{\sigma}$ and hence $\bar{c}_{\sigma}(\bar{\lambda})=c_{\tilde{\sigma}}(\lambda)$. Now we define the wave packet transform on ${\cal H}_0$. Later we will extend it by continuity to a Hilbert space closure of ${\cal H}_0$. \begin{ddd} The wave packet transform is the map $E:{\cal H}_0\rightarrow C^\infty(Y,V(\gamma))$ given by $$E(\phi) := \int_0^\infty E(\imath\mu,\phi_{\imath\mu},T(\imath\mu))\:d\mu\ .$$ The section $E(\phi)$, $\phi\in{\cal H}_0$, is called a wave packet (of Eisenstein series). \end{ddd} \begin{lem}\label{swa} If $\phi\in {\cal H}_0$, then $E(\phi)\in S(Y,V_Y(\gamma))$. \end{lem} {\it Proof.$\:\:\:\:$} We reduce the proof of this lemma to the case where $\Gamma$ is trivial. In this case the assertion is well known \cite{arthur75}. Let $W\subset X\cup\Omega$ be a compact subset such that $\Gamma W=X\cup\Omega$. Let $\partial W:=W\cap\Omega$ and let $\chi\in C_c^\infty(\Omega)$ be a cut-off function such that $\chi_{|\partial W}=1$. Let $\tilde{\phi}_{\imath\mu}=ext\:\phi_{\imath\mu}$ and set $\psi_{\imath\mu}=\chi \tilde{\phi}_{\imath\mu}$, $\xi_{\imath\mu}=(1-\chi)\tilde{\phi}_{\imath\mu}$. Then ${\bf a}^*_+\ni\mu\mapsto \psi_{\imath\mu}\in C^\infty(\partial X,V(\sigma_{\imath\mu}))$ is a smooth family with compact support. It was shown in \cite{arthur75} that $$P(\psi):=\int P^T_{\imath\mu}(\psi_{\imath\mu}) d\mu\in S(X,V(\gamma))\ .$$ Since $E(\phi)=P(\psi)+P(\xi)$, it remains to show that $q_{W,A,N}(P(\xi))<\infty$ for all $A\in {\cal U}({\bf g})$, $N\in{\bf N}$, where $q_{W,A,N}$ is one of the seminorms (\ref{swart}) characterizing the Schwartz space. By the $G$-equivariance of the Poisson transform we have $$L_AP(\xi)=\int P^T_{\imath\mu}(\pi^{\sigma,\imath\mu}(A) \xi_{\imath\mu}) d\mu\ .$$ Note that $\xi_{|\partial W} =0$. The expansion of the Poisson transform obtained in Lemma \ref{eee} leads to the following decomposition into a leading and a remainder term. For $k\in \partial WM$ \begin{eqnarray*} P^T_{\imath\mu}(\pi^{\sigma,\imath\mu}(A) \xi_{\imath\mu})(ka)&=& a^{-(\imath\mu+\rho)}\gamma(w)T (\hat{J}_{\imath\mu}(\pi^{\sigma,\imath\mu}(A) \xi_{\imath\mu}))(k)\\ &&+ a^{-(\imath\mu+\rho+\alpha_1)} p_1(a,k,\pi^{\sigma,\imath\mu}(A) \xi_{\imath\mu})\ , \end{eqnarray*} where $p_1(a,k,f)$ is uniformly bounded as $ka\in W$ in terms of the distribution $f$. Thus for all $N$ we have $$\{W\ni ka\mapsto |\log(a)|^N \int a^{-(\imath\mu+\rho+\alpha_1)} p_1(a,k,\pi^{\sigma,\imath\mu}(A) \xi_{\imath\mu}) d\mu\}\in L^2(W,V(\gamma))\ .$$ This shows that the remainder term leads to something satisfying the estimates of the Schwartz space. We now analyse the leading term. Since the family $\xi_{\imath\mu}$ has compact support with respect to $\mu$ and $\hat{J}_{\imath\mu}(\pi^{\sigma,\imath\mu}(A) \xi_{\imath\mu})$ is smooth in $(\mu,k)$ by Lemma \ref{off} we obtain for all $N\in{\bf N}_0$ $$|\int a^{-(\imath\mu+\rho)} \gamma(w) T (\hat{J}_{\imath\mu}(\pi^{\sigma,\imath\mu}(A) \xi_{\imath\mu}))(k) d\mu|\le C_N (1+|\log(a)|)^{-N} a^{-\rho},\quad \forall ka\in W$$ for each $N\in{\bf N}$. Thus the leading terms satisfies the Schwartz space estimates, too. This proves the lemma. \hspace*{\fill}$\Box$ \\[0.5cm]\noindent Let $\mu\not=0$, ${\rm Re }(\mu)=0$, and $\psi\in C^{-\infty}(B,V_B(\tilde{\sigma}_{-\imath\mu}))$. Set $\tilde{\psi}:=E(-\imath\mu,\psi,\tilde{T}(-\imath\mu))$. \begin{lem}\label{tttq} $\tilde{\psi}\in S(Y,V_Y(\gamma))^\prime$. \end{lem} {\it Proof.$\:\:\:\:$} We have $ext (\psi)\in C^{-\infty}(\partial X,V(\tilde{\sigma}_{-\imath\mu}))$. It known by \cite{arthur75} that $P^{\tilde{T}}_{-\imath\mu}(ext(\psi))\in S(X,V(\gamma))^\prime$. This implies the lemma. \hspace*{\fill}$\Box$ \\[0.5cm]\noindent By Lemmas \ref{swa} and \ref{tttq} the pairing $\langle E(\phi),\tilde{\psi}\rangle$ between the wave packet $E(\phi)$ and the generalized eigensection $\tilde{\psi}$ is well defined. The following proposition gives an explicit formula for this pairing. \begin{prop}\label{scalar} We have $\langle E(\phi),\tilde{\psi}\rangle=\pi\langle \phi_{\imath\mu},\psi\rangle$. \end{prop} {\it Proof.$\:\:\:\:$} Let $\psi_n\in C^{\infty}(B,V_B(\tilde{\sigma}_{-\imath\mu}))$ be a sequence approximating the distribution $\psi$. Then by \cite{arthur75} and the continuity of $ext$ we have $$E(-\imath\mu,\psi_n,\tilde{T}(-\imath\mu))\to E(-\imath\mu,\psi,\tilde{T}(-\imath\mu))$$ in $S(Y,V_Y(\gamma))^\prime$ as $n\to \infty$. Thus the proposition is a consequence of the following the special case. \begin{lem} Let $\psi\in C^{\infty}(B,V_B(\tilde{\sigma}_{-\imath\mu}))$, then $\langle E(\phi),\tilde{\psi}\rangle=\pi\langle \phi_{\imath\mu},\psi\rangle$. \end{lem} {\it Proof.$\:\:\:\:$} Let $W\subset\Omega$ be compact. The following asymptotic expansions hold uniformly for $k\in WM$ and $a\in A_+$ large: \begin{eqnarray} E(\imath \mu,\phi_{\imath\mu},T(\imath\mu))(ka)& =& a^{\imath\mu -\rho}\frac{c_\gamma(\imath\mu)}{c_\sigma(\imath\mu)} T ext(\phi_{\imath\mu})(k)\nonumber\\ &&+a^{-\imath\mu-\rho}\gamma(w)T \frac{c_\sigma(-\imath\mu)}{c_\sigma(\imath\mu)} ext(S_{\imath\mu}\phi_{ \imath\mu})(k) + O(a^{-\rho-\epsilon})\label{w1w}\\ E(-\imath\mu,\psi_{-\imath\mu},\tilde{T}(-\imath\mu))(ka) & = & a^{-\imath\mu -\rho} \frac{c_{\tilde{\gamma}}(-\imath\mu)}{ c_{\tilde{\sigma}} (-\imath\mu)} \tilde{T} ext(\psi_{-\imath\mu})(k)\nonumber \\ &&+a^{ \imath\mu-\rho}\tilde{\gamma}(w)\tilde{T} \frac{c_{\tilde{\sigma}}(\imath\mu)}{ c_{\tilde{\sigma}} (-\imath\mu)} ext(S_{-\imath\mu}\psi_{ -\imath\mu})(k) + O(a^{-\rho-\epsilon})\ .\nonumber \end{eqnarray} These expansions are immediate consequences of the asymptotic expansion of the Poisson transform of smooth sections and can be differentiated with respect to $a$ and differentiated and integrated with respect to $\mu$. In order to read these formulas appropriately identify sections of $V_Y(\gamma)$ with $\Gamma$-invariant functions on $G$ with values in $V_\gamma$, sections of $V(\sigma_{\imath\mu})$ with functions on $G$ with values in $V_\sigma$, etc., as usual. Let $\chi$ be a cut-off function as constructed in Lemma \ref{lll} and $B_R$ the ball of radius $R$ around the origin of $X$. If we define $A_{\imath\mu}:=-\Omega_G+\chi_{\mu_\sigma+\rho_m-\imath\mu}(\Omega_G)$, then $A_{\imath\mu}E(\imath \mu,\phi_{\imath\mu},T(\imath\mu))=0$ and $A_{-\imath\mu} E(-\imath\mu,\psi_{-\imath\mu},\tilde{T}(-\imath\mu))=0$. We start with the following identity \begin{eqnarray*}0&=& \langle\chi A_{ \imath\lambda}E(\imath\lambda,\phi_{\imath\lambda},T(\imath\lambda)), E(-\imath\mu,\psi_{-\imath\mu},\tilde{T}(-\imath\mu))\rangle_{L^2(B_R)}\\&& - \langle\chi E(\imath \lambda,\phi_{\imath\lambda},T(\imath\lambda)), A_{-\imath\mu} E(-\imath\mu,\psi_{-\imath\mu},\tilde{T}(-\imath\mu)) \rangle_{L^2(B_R)}\ .\end{eqnarray*} By partial integration as in the proof of Proposition \ref{green} we obtain \begin{eqnarray} \lefteqn{(\mu^2-\lambda^2)\langle \chi E(\imath\lambda,\phi_{\imath\lambda},T(\imath\lambda)), E(-\imath\mu,\psi_{-\imath\mu},\tilde{T}(-\imath\mu))\rangle_{L^2(B_R)}}\hspace{3cm}\\ &=&\langle \chi \nabla_n E(\imath\lambda,\phi_{\imath\lambda},T(\imath\lambda)), E(-\imath\mu,\psi_{-\imath\mu},\tilde{T}(-\imath\mu))\rangle_{L^2(\partial B_R)}\label{w2e}\\ &&-\langle \chi E(\imath\lambda,\phi_{\imath\lambda},T(\imath\lambda)), \nabla_n E(-\imath\mu,\psi_{-\imath\mu},\tilde{T}(-\imath\mu))\rangle_{L^2(\partial B_R)}\label{w3e}\\ &&-\langle [A_0,\chi] E(\imath\lambda,\phi_{\imath\lambda},T(\imath\lambda)), E(-\imath\mu,\psi_{-\imath\mu},\tilde{T}(-\imath\mu))\rangle_{L^2(B_R)}\ .\label{w4e} \end{eqnarray} We insert the asymptotic expansions (\ref{w1w}) which hold on the support of $\chi$. We obtain with $a_R:={\rm e}^{R}$ \begin{eqnarray*} (\ref{w2e})+(\ref{w3e}) &=& \imath(\lambda+\mu) a_R^{\imath (\lambda -\mu)}\langle\chi \frac{c_\gamma(\imath\lambda)}{c_\sigma(\imath\lambda)} T ext(\phi_{\imath\lambda}),\frac{c_{\tilde{\gamma}}(-\imath\mu)}{ c_{\tilde{\sigma}} (-\imath\mu)} \tilde{T} ext(\psi_{-\imath\mu})\rangle \\ &&+\imath (\lambda-\mu) a_R^{\imath(\lambda+\mu)} \langle\chi \frac{c_\gamma(\imath\lambda)}{c_\sigma(\imath\lambda)} T ext(\phi_{\imath\lambda}), \tilde{\gamma}(w) \tilde{T} \frac{c_{\tilde{\sigma}}(\imath\mu) }{c_{\tilde{\sigma}} (-\imath\mu)} ext(S_{-\imath\mu}\psi_{-\imath\mu})\rangle\\ &&+\imath(-\lambda+\mu) a_R^{\imath(-\lambda-\mu)} \langle\chi \gamma(w) T \frac{c_\sigma(-\imath\lambda)}{ c_\sigma(\imath\lambda) } ext(S_{\imath\lambda} \phi_{\imath\lambda}), \frac{c_{\tilde{\gamma}}(-\imath\mu)}{ c_{\tilde{\sigma}} (-\imath\mu)}\tilde{T} ext(\psi_{-\imath\mu})\rangle\\ &&+\imath(-\lambda-\mu) a_R^{\imath(-\lambda+\mu)} \langle\chi \frac{c_\sigma(-\imath\lambda) c_{\tilde{\sigma}}(\imath\mu)}{c_\sigma(\imath\lambda )c_{\tilde{\sigma}}(-\imath\mu)} ext(S_{\imath\lambda}\phi_{\imath\lambda}), ext(S_{-\imath\mu}\psi_{-\imath\mu})\rangle\ .\\ &&o(1)\ . \end{eqnarray*} The pairings on the right-hand side are defined using the canonical $K$-equivariant identification of the bundles $V(\sigma_\lambda)$ with $V(\sigma_0)$. We combine the remainder $o(1)$ and the term (\ref{w4e}) to $F(\lambda,\mu,R)$. Then we can write \begin{eqnarray} \lefteqn{ \langle \chi E(\imath\lambda,\phi_{\imath\lambda},T(\imath\lambda)), E(-\imath\mu,\psi_{-\imath\mu},\tilde{T}(-\imath\mu))\rangle_{L^2(B_R)}}\hspace{2cm}\label{e34r}\\ &=& \imath \frac{a_R^{\imath (\lambda -\mu)}}{ -\lambda+\mu } \langle\chi \frac{c_\gamma(\imath\lambda)}{c_\sigma(\imath\lambda)} T ext(\phi_{\imath\lambda}),\frac{c_{\tilde{\gamma}}(-\imath\mu)}{ c_{\tilde{\sigma}} (-\imath\mu)} \tilde{T} ext(\psi_{-\imath\mu})\rangle\label{sing1} \\ &&-\imath \frac{a_R^{\imath(\lambda+\mu)}}{ \lambda+\mu} \langle\chi ext(\phi_{\imath\lambda}), T^*\frac{c^*_\gamma(\imath\mu)}{c_\sigma(\imath\mu)}\tilde{\gamma}(w) \tilde{T} \frac{c_{\tilde{\sigma}}(\imath\mu)}{ c_{\tilde{\sigma}} (-\imath\mu) } ext(S_{-\imath\mu}\psi_{-\imath\mu})\rangle\nonumber\\ &&+\imath \frac{a_R^{\imath(-\lambda-\mu)}}{\lambda+\mu} \langle\chi \tilde{T}^*\frac{c^*_{\tilde{\gamma}}(-\imath\mu)}{ c_{\tilde{\sigma}} (-\imath\mu)}\gamma(w) T \frac{c_\sigma(-\imath\lambda)}{ c_\sigma(\imath\lambda) }ext(S_{\imath\lambda} \phi_{\imath\lambda}), ext(\psi_{-\imath\mu})\rangle\nonumber\\ &&+\imath \frac{a_R^{\imath(-\lambda+\mu)}}{\lambda-\mu} \langle\chi \frac{c_\sigma(-\imath\lambda)c_{\tilde{\sigma}}(\imath\mu)}{ c_\sigma(\imath\lambda )c_{\tilde{\sigma}}(-\imath\mu)} ext(S_{\imath\lambda}\phi_{\imath\lambda}), ext(S_{-\imath\mu}\psi_{-\imath\mu})\rangle\label{sing2}\\ &&+\frac{F(\lambda,\mu,R)}{\mu^2-\lambda^2}\ . \end{eqnarray} Since $S_{\imath\mu}$ is unitary, $$\frac{c_\sigma(-\imath\lambda)c_{\tilde{\sigma}}(\imath\lambda)}{ c_\sigma(\imath\lambda )c_{\tilde{\sigma}}(-\imath\lambda)} =1\ ,$$ and \begin{equation}\label{wonderid}P_{\sigma}(\imath\mu)^{-1}{\mbox{\rm id}}_{V_{\tilde{\gamma}}(\tilde{\sigma})}= c_\sigma(\imath\mu)c_{\tilde{\sigma}}(-\imath\mu){\mbox{\rm id}}_{V_{\tilde{\gamma}}(\tilde{\sigma})}=T^* c_\gamma(\imath\mu)^*c_{\tilde{\gamma}}(-\imath\mu)\tilde{T}\ ,\end{equation} the singularities of the terms (\ref{sing1}) and (\ref{sing2}) at $\mu=\lambda$ cancel and $\frac{F(\lambda,\mu,R)}{\mu^2-\lambda^2}$ is smooth at $\mu=\lambda$. Moreover $F(\lambda,\mu,R)\to 0$ as $R\to \infty$ such that the $C^1$-norm with respect to $\lambda$ remains bounded. By Lebesgue's theorem about dominated convergence we obtain $$\lim_{R\to\infty} \int_0^\infty \frac{F(\lambda,\mu,R)}{\mu^2-\lambda^2} d\lambda = 0\ .$$ By the Lemma of Riemann-Lebesgue \begin{eqnarray*} \lim_{R\to\infty}\int_0^\infty \imath \frac{a_R^{\imath(-\lambda-\mu)}}{\lambda+\mu} \langle\chi \tilde{T}^*\frac{c^*_\gamma(\imath\mu)}{c_\sigma(\imath\mu)}\gamma(w) T \frac{c_\sigma(-\imath\lambda)}{ c_\sigma(\imath\lambda)} ext(S_{\imath\lambda} \phi_{\imath\lambda}), ext(\psi_{-\imath\mu})\rangle d\lambda &=& 0 \\ \lim_{ R\to\infty}\int_0^\infty \imath \frac{a_R^{\imath(\lambda+\mu)}}{ \lambda+\mu} \langle\chi ext(\phi_{\imath\lambda}), T^*\frac{c^*_{\tilde{\gamma}}(-\imath\mu)}{ c_{\tilde{\sigma}} (-\imath\mu)}\tilde{\gamma}(w) \tilde{T}\frac{c_{\tilde{\sigma}}(\imath\mu)}{ c_{\tilde{\sigma}} (-\imath\mu)} ext(S_{-\imath\mu}\psi_{-\imath\mu})\rangle d\lambda &=&0\ . \end{eqnarray*} We set $s:=\lambda-\mu$. We regroup the remaining terms of (\ref{e34r}) to \begin{eqnarray*} \lefteqn{\frac{a_R^{\imath s}-a_R^{-\imath s}}{\imath s} \langle\chi \frac{c_\gamma(\imath\lambda)}{c_\sigma(\imath\lambda)} T ext(\phi_{\imath\lambda}),\frac{c_{\tilde{\gamma}}(-\imath\mu)}{ c_{\tilde{\sigma}} (-\imath\mu)} \tilde{T} ext(\psi_{-\imath\mu})\rangle}\hspace{0cm}\\ &&+a_R^{-\imath s } \frac{\langle\chi \frac{c_\gamma(\imath\lambda)}{c_\sigma(\imath\lambda)} T ext(\phi_{\imath\lambda}),\frac{c_{\tilde{\gamma}}(-\imath\mu)}{ c_{\tilde{\sigma}} (-\imath\mu)} \tilde{T} ext(\psi_{-\imath\mu})\rangle- \langle\chi \frac{c_\sigma(-\imath\lambda)c_{\tilde{\sigma}}(\imath\mu) }{c_\sigma(\imath\lambda )c_{\tilde{\sigma}}(-\imath\mu)} ext(S_{\imath\lambda}\phi_{\imath\lambda}), ext(S_{-\imath\mu}\psi_{-\imath\mu})\rangle}{\imath s} \end{eqnarray*} If we integrate the second term with respect to $s$ and perform the limit $R\to\infty$, then the result vanishes by the Riemann-Lebesgue lemma. Using the identity of distributions $\lim_{r\to\infty}\frac{\sin(rs)}{s }=\pi\delta_0(s)$ and (\ref{wonderid}) the first term gives $$ \lim_{R\to\infty} \int_{-\infty}^\infty \frac{a_R^{\imath s}-a_R^{-\imath s}}{2 \imath s} \langle\chi \frac{c_\gamma(\imath\lambda)}{c_\sigma(\imath\lambda)} T ext(\phi_{\imath\lambda}), \frac{c_{\tilde{\gamma}}(-\imath\mu)}{ c_{\tilde{\sigma}} (-\imath\mu)} \tilde{T} ext(\psi_{-\imath\mu})\rangle ds = \pi \langle \phi_{\imath\mu} , \psi \rangle\ . $$ The limit as $R\to\infty$ of the integral of left-hand side of (\ref{e34r}) with respect to $\lambda$ is equal to $\langle E(\phi),\tilde{\psi}\rangle$. This proves the lemma and thus finishes the proof of the proposition. \hspace*{\fill}$\Box$ \\[0.5cm]\noindent In order to deal with Hilbert spaces we go over to employ sesquilinear pairings which will be denoted by $(.,.)$. The unitary structure of $\sigma$ induces a conjugate linear isomorphism of $\tilde{\sigma}_{-\imath\mu}$ with $\sigma_{\imath\mu}$. Analogously the unitary structure of $\gamma$ induces a conjugate linear isomorphism of $\tilde{\gamma}$ with $\gamma$. We choose $T$ to be unitary, then $\tilde{T}$ corresponds to $T$ under the above identifications. Moreover, if $\psi\in C^{-\infty}(B,V_B(\tilde{\sigma}_{-\imath\mu}))$ corresponds to $\bar{\psi}\in C^{-\infty}(B,V_B(\sigma_{\imath\mu}))$, then $\tilde{\psi}$ corresponds to $\tilde{\bar{\psi}}:= E(\imath\mu,\bar{\psi},T(\imath\mu))$ under the isomorphisms above. Using the sesquilinear pairings on $\gamma$ and $\sigma_{\imath\mu}$ we can rewrite the result of Proposition \ref{scalar} as \begin{equation}\label{sscar}( E(\phi),\tilde{\bar{\psi}} )=\pi( \phi_{\imath\mu},\bar{\psi})\ .\end{equation} We define a scalar product on ${\cal H}_0$ by $$( \phi,\psi):=\pi \int_0^\infty ( \phi_{\imath\mu},\psi_{\imath\mu}) d\mu$$ and let ${\cal H}$ be the corresponding Hilbert space closure of ${\cal H}_0$. The following corollary is a consequence of Proposition \ref{scalar}. \begin{kor}\label{hermit} The wave packet transform extends by continuity to an isometric embedding $E:{\cal H}\hookrightarrow L^2(Y,V_Y(\gamma))$. \end{kor} If $A\in{\cal Z}$ and $\phi\in{\cal H}_0$, then we have $AE(\phi)=E(\psi)$ with $\psi_{\imath\mu}=\chi_{\mu_\sigma+\rho_m-\imath\mu}(A)\phi_{\imath\mu}$. Choose an orthogonal decomposition $(V_\gamma)_{|M}=\oplus_iV_{\sigma_i}$ and let $T_i\in {\mbox{\rm Hom}}_M(V_{\sigma_i},V_\gamma)$ be the corresponding unitary embeddings. Here some of the $\sigma_i$ may be equivalent. Let ${\cal H}(i)$ be the Hilbert space corresponding to $\sigma_i$ and $E_i$ the corresponding wave packet transform. It is easy to modify the proof of Proposition \ref{scalar} in order to show that the ranges of the $E_i$ are pairwise orthogonal. We define the unitary embedding $$E_\gamma:=\oplus_i E_i: {\cal H}(\gamma):=\bigoplus_i{\cal H}(i)\hookrightarrow L^2(Y,V_Y(\gamma))\ .$$ Then $E_\gamma$ represents an absolute-continuous subspace $L^2(Y,V_Y(\gamma))_c\subset L^2(Y,V_Y(\gamma))$ with respect to ${\cal Z}$. We now prove that the orthogonal complement of the range of $E_\gamma$ is the discrete subspace and the corresponding characters belong to $PS$. The abstract spectral decomposition of $L^2(Y,V_Y(\gamma))$ with respect to the commutative algebra ${\cal Z}$ provides an unitary equivalence $$\alpha:L^2(Y,V_Y(\gamma))\cong H:= \int_{{\bf h}_{\bf C}^*/W} H_\lambda \kappa(d\lambda)\ ,$$ where the Hilbert space $H_\lambda$ is a ${\cal Z}$-module on which ${\cal Z}$ acts by $\chi_\lambda$. A part of the structure of the direct integral is that $H$ is a space of sections $ {\bf h}_{\bf C}^*/W\ni\lambda\mapsto \psi_\lambda\in H_\lambda$ such that ${\rm clo}\{\psi_\lambda | \psi\in H\} = H_\lambda$ ($\kappa$-almost everywhere), and such that the scalar products ${\bf h}_{\bf C}^*/W\ni \lambda\mapsto(\psi_\lambda,\psi_\lambda^\prime)\in{\bf C}$, $\psi,\psi^\prime\in H$, are measurable functions. Then scalar product on $H$ is given by $$(\psi,\psi^\prime)= \int_{{\bf h}_{\bf C}^*/W}(\psi_\lambda,\psi_\lambda^\prime)\kappa(d\lambda)\ .$$ Fix a base $\{X_i\}$ of ${\bf g}$ and let $I_N$, $N\in{\bf N}_0$, denote the set of all multiindices $i=(i_1,\dots,i_{\dim({\bf g})})$, $|i|\le N$. Let $\chi$ be the cut-off function constructed in Lemma \ref{lll}. For $f\in S(Y,V_Y(\gamma))$ we define $$\|f\|^2_N:=\sum_{i\in I_N}\int_G \chi(gk) |\log(a(g))|^N |f(X_i g)|^2 dg \ .$$ Note that if $f\in S(Y,V_Y(\gamma))$, then $\|f\|_N<\infty$. By $S^N(Y,V_Y(\gamma))$ we denote the closure of the Schwartz space $S(Y,V_Y(\gamma))$ with respect to $\|.\|_N$. Then $S^N(Y,V_Y(\gamma))$ is a Hilbert space contained in $L^2(Y,V_Y(\gamma))$. \begin{lem}\label{komppp} If $N$ is sufficiently large, then the inclusion $$S^N(Y,V_Y(\gamma))\hookrightarrow L^2(Y,V_Y(\gamma))$$ is Hilbert-Schmidt. \end{lem} {\it Proof.$\:\:\:\:$} This follows from the results of \cite{bernstein88}. In order to provide some details we employ the notions "space of polynomial growth" and "comparable scale functions" introduced in \cite{bernstein88}. Fix some base point $y\in Y$. Let the scale function $r:\Gamma\backslash G\rightarrow {\bf R}^+$ be given by $r(\Gamma g)={\rm dist}_Y(y,\Gamma g K)$, where ${\rm dist}_Y$ denotes the Riemannian distance in $Y$. Then $\Gamma\backslash G$ is a space of polynomial growth with respect to $r$, i.e., if $A\subset G$ is a compact neighbourhood of the identity, then there exist constants $d\ge 0$, $C>0$ such that for any $R>0$ there exists a set of $\le C(1+R)^d$ points $x_i$ of $\Gamma\backslash G$ such that $\{r(x)<R\}\subset \cup_i x_i A$. This follows essentially from the fact that $G$ itself is of polynomial growth and that ${\rm dist}_X(x,.)$ and ${\rm dist}_Y(y,.)$ are comparable when restricted to a compact fundamental domain $F\subset X\cup\Omega$ containing the lift $x$ of $y$. Note that $g\mapsto r(\Gamma g)$ and $g\mapsto a(g)$ are comparable on this fundamental domain, too. If we choose $N>\max\{d,\dim({\bf g})\}$, then the assertion of the lemma is proved by the Proposition of \cite{bernstein88}, Sec. 3.4. \hspace*{\fill}$\Box$ \\[0.5cm]\noindent In the following we choose $N$ suffiently large. It follows by a theorem of Gelfand/Kostyuchenko (see \cite{bernstein88}) that the composition $$S^N(Y,V_Y(\gamma))\hookrightarrow L^2(Y,V_Y(\gamma))\rightarrow \int_{{\bf h}_{\bf C}^*/W} H_\lambda \kappa(d\lambda)$$ is pointwise defined, i.e., there exists a collection of continuous maps $$\alpha_\lambda:S^N(Y,V_Y(\gamma))\rightarrow H_\lambda,\quad \lambda\in{\bf h}^*_{\bf C}/W$$ such that for $\phi\in S^N(Y,V_Y(\gamma))$ we have $\alpha(\phi)_\lambda=\alpha_\lambda(\phi)$. Let $S^N(Y,V_Y(\gamma))^*$ denote the Hermitean dual of $S^N(Y,V_Y(\gamma))$. Then we have inclusions $$S(Y,V_Y(\gamma))\hookrightarrow S^N(Y,V_Y(\gamma))\hookrightarrow L^2(Y,V_Y(\gamma))\hookrightarrow S^N(Y,V_Y(\gamma))^*\hookrightarrow S(Y,V_Y(\gamma))^*\ .$$ By changing $\alpha_\lambda$ on a set of $\lambda$'s of measure zero (mod $\kappa$) we can assume that for all $\lambda\in{\bf h}_{\bf C}^*/W$ the map $$\alpha_\lambda:S(Y,V_Y(\gamma))\rightarrow H_\lambda$$ is a morphism of ${\cal Z}$-modules. Let $$\beta_\lambda:H_\lambda\rightarrow S^N(Y,V_Y(\gamma))^*$$ denote the adjoint of $\alpha_\lambda$. Since $$\beta_\lambda:H_\lambda\rightarrow S(Y,V_Y(\gamma))^*$$ is a morphism of ${\cal Z}$-modules we see that $\beta_\lambda(H_\lambda)$ consists of tempered eigensections of ${\cal Z}$ corresponding to the character $\chi_\lambda$. \begin{prop}\label{ortho} Let $\psi\in L^2(Y,V_Y(\gamma))$ be represented by $\alpha(\psi)$ such that $\alpha(\psi)_\lambda=0$ for all $\lambda\in PS$. Assume further that $(E_\gamma(\phi),\psi)=0$ for all wave packets $E_\gamma(\phi)$, $\phi=\oplus_i\phi_i\in \oplus_i{\cal H}_0(i) $. Then $\psi=0$. \end{prop} {\it Proof.$\:\:\:\:$} Let $\phi=\oplus\phi_i\in \oplus {\cal H}_0(i)$. Then we can write \begin{eqnarray*} 0&=&( E_\gamma(\phi),\psi)\\ &=&\int_{{\bf h}_{\bf C}^*/W}(\alpha_\lambda(E_\gamma(\phi)),\alpha(\psi)_\lambda) \kappa(d\lambda)\\ &=&\int_{{\bf h}_{\bf C}^*/W} (E_\gamma(\phi),\beta_\lambda\alpha(\psi)_\lambda)\kappa(d\lambda)\ . \end{eqnarray*} We claim that for $\lambda\not\in PS$ we can write $$\beta_\lambda\alpha(\psi)_\lambda=\sum_i E_\gamma(\imath\mu_i,\psi_{i,\imath\mu_i},T_i(\imath\mu_i))\ ,$$ where $\psi_{i,\imath\mu_i}\in C^{-\infty}(B,V_B(\sigma_{i,\imath\mu_i}))$, ${\rm Im}(\mu_i)=0$ if $\psi_{i,\imath\mu_i}\not=0$, and $\lambda=\mu_{\sigma_i}+\rho_m-\imath\mu_i$. In fact, by Proposition \ref{gener1} we have $\beta_\lambda\alpha(\psi)_\lambda=\sum_i P_{\mu_i}^{T_i}(\tilde{\psi}_{i,\imath\mu_i})$, where $\tilde{\psi}_{i,\imath\mu_i}\in {}^\Gamma C^{-\infty}(\partial X,V(\sigma_{i,\imath\mu_i}))$. If ${\rm Im}(\mu_i)\not=0$ and $\tilde{\psi}_{i,\imath\mu_i}\not=0$, then ${\mbox{\rm supp}}(\tilde{\psi}_i)\in \Lambda$ by Lemma \ref{wo}. But then $\lambda \in PS_d$ and this case was excluded. By the functional equation of the Eisenstein series (Corollary \ref{funeq}) and since ${\bf a}^*\ni\mu_i\not=0$ (because of $\lambda\not\in PS$) we can assume that ${\rm Re }(\mu_i)>0$ for all relevant $i$. By Proposition \ref{upperbound} we have $\tilde{\psi}_{i,\imath\mu_i}=ext (\Psi_{i,\imath\mu_i})$ with $\Psi_{i,\imath\mu_i}=res (\tilde{\psi}_{i,\imath\mu_i})$. Putting $\psi_{i,\imath\mu_i}=c_\sigma(\imath\mu)\Psi_{i,\imath\mu_i}$ we obtain the claim. We consider $\mu_i$ as a function of $\lambda$. Then using (\ref{sscar}) we obtain \begin{eqnarray*} 0&=&\sum_i\int_{{\bf h}_{\bf C}^*/W\setminus PS} ( E_\gamma(\phi),E(\imath\mu_i,\psi_{i,\imath\mu_i},T_i(\imath\mu_i))) \kappa(d\lambda)\\ &=&\pi\sum_i \int_{{\bf h}_{\bf C}^*/W\setminus PS} (\phi_{i,\imath\mu_i},\psi_{i,\imath\mu_i}) \kappa(d\lambda)\ . \end{eqnarray*} Let $f_i\in C_c^\infty({\bf h}_{\bf C}^*/W\setminus PS)$. Then $\{ (0,\infty) \ni \mu\rightarrow f_i(\lambda(\mu))\phi_{i,\imath\mu}\}\in{\cal H}_0(i)$ and thus \begin{equation}\label{opop}0=\sum_i \int_0^\infty f_i(\mu_i) (\phi_{i,\imath\mu_i},\psi_{i,\imath\mu_i}) \kappa(d\lambda)\ .\end{equation} We conclude that $( \phi_{i,\imath\mu_i},\psi_{i,\imath\mu_i}) =0$ for almost all $\lambda$ (mod $\kappa$). We now trivialize the family of bundles by identifying $V_B(\sigma_{i,\imath\mu})$ with $V_B(\sigma_{i,0})$ in some holomorphic manner. We choose a countable dense set $\{\phi_j\}\subset C^\infty(B,V_B(\sigma_{i,0}))$. Viewing $\mu\mapsto \psi_{i,\imath\mu_i}$ as a family of sections in $C^{-\infty}(B,V_B(\sigma_{i,0}))$ we form $B_j:=\{\lambda|(\phi_j,\psi_{i,\imath\mu_i})\not=0\}\subset {\bf h}^*_{\bf C}/W$. Then $\kappa(B_j)=0$. Moreover let $U:=\cup_j B_j$. Then $\kappa(U)=0$ and we have $(\phi_j,\psi_{i,\imath\mu_i})=0$ for all $\lambda\in {\bf h}_{\bf C}^*/W_+\setminus U$ and all $j$. Thus $\psi_{i,\imath\mu_i}=0$ for $\lambda\in {\bf h}_{\bf C}^*/W_+\setminus U$. Hence $\alpha(\psi)_\lambda=0$ for almost all $\lambda$ (mod $\kappa$). Hence $\psi=0$. \hspace*{\fill}$\Box$ \\[0.5cm]\noindent The following theorem is the immediate consequence of Proposition \ref{ortho}. \begin{theorem}\label{contsp} The wave packet transform $E_\gamma$ is an unitary equivalence of ${\cal H}$ with the absolute-continuous subspace of $L^2(Y,V_Y(\gamma))$. The orthogonal complement of the absolute-continuous subspace is the discrete subspace $L^2(Y,V_Y(\gamma))$ and the corresponding eigencharacters belong to $PS$. \end{theorem} \section{The discrete spectrum} In Section \ref{wxa} we obtained a complete description of the continuous subspace $L^2(Y,V_Y(\gamma))$ in terms of the wave packet transform. In the present section we study the othogonal complement $L^2(Y,V_Y(\gamma))_d$ of the continuous subspace. The discrete subspace decomposes further into a cuspidal, residual, and scattering component (see Definition \ref{t6r}). We show that the residual and the scattering component are finite-dimensional and that the cuspidal component is either trivial or infinite-dimensional. The notions of the cuspidal and the residual components are similar to the corresponding notions known in the finite volume case. The appearence of the scattering component is a new phenomenon which does not occur in the finite volume case. We give examples where the scattering component is non-trivial. In order to study the eigenspaces of ${\cal Z}$ on $L^2(Y,V_Y(\gamma))$ we decompose them further with respect to the full algebra of invariant differential operators ${\cal D}_\gamma$ on $V(\gamma)$. We first recall some facts concerning ${\cal D}_\gamma$ (see \cite{olbrichdiss}). The algebra ${\cal D}_\gamma$ is in general a non-commutative finite extension of ${\cal Z}_\gamma$. The right action of ${\cal U}({\bf g})^K$ on $C^\infty(X,V(\gamma))$ induces a surjective homomorphism ${\cal U}({\bf g})^K\rightarrow {\cal D}_\gamma$. Hence any representation of $D_\gamma$ can be lifted to a representation of ${\cal U}({\bf g})^K$. For $\sigma\in\hat{M}$ and $\lambda\in{\aaaa_\C^\ast}$ there is a representation of $\chi_{\sigma,\lambda}$ of ${\cal U}({\bf g})^K$ into ${\mbox{\rm End}}_M(V_\gamma(\sigma))$ which descends to ${\cal D}_\gamma$ such that its restriction to ${\cal Z}_\gamma$ induces $\chi_{\mu_\sigma+\rho_m-\lambda}$. Here $V_\gamma(\sigma)$ denote the $\sigma$-isotypic component of $V_\gamma$. The representation $\chi_{\sigma,\lambda}$ is characterized by \begin{equation}\label{nearby}D\circ P^T_\lambda=P^{\chi_{\sigma,\lambda}(D)\circ T}_\lambda,\quad T\in {\mbox{\rm Hom}}_M(V_\sigma,V_\gamma),\:\: D\in D_\gamma\ ,\end{equation} and where $P^T_\lambda$ denotes the Poisson transform. To be more precise, the representations $\chi_{\sigma,\lambda}$ depend holomorphically on $\lambda$ and (\ref{nearby}) is an identity between holomorphic families of maps. By ${\cal E}_{\sigma,\lambda}(X,V(\gamma))$ we denote the space of all $f\in C^\infty(X,V(\gamma))$ which under ${\cal D}_\gamma$ generate a quotient of the representation $\chi_{\sigma,\lambda}$. Let ${\cal E}_{\sigma,\lambda}(Y,V_Y(\gamma))$ be the subspace of ${\cal E}_{\sigma,\lambda}(X,V(\gamma))$ of $\Gamma$-invariant sections. Then ${\cal Z}$ acts on ${\cal E}_{\sigma,\lambda}(X,V(\gamma))$ and ${\cal E}_{\sigma,\lambda}(Y,V_Y(\gamma))$ by the character $\chi_{\mu_\sigma+\rho_m-\lambda}$. Define $PS_{res}(\sigma)\subset {\aaaa_\C^\ast}$ by $PS_{res}(\sigma):=\{\mu \:|\:{\rm Re }(\mu)>0,\: {}^\Gamma C^{-\infty}(\Lambda,V(\sigma_\mu))\not= 0\}$. By Proposition \ref{upperbound}, $3.$ we have $PS_{res}(\sigma)\subset (0,\rho)$. \begin{prop}\label{ddsp} \begin{enumerate} \item $PS_{res}(\sigma)$ is a finite set. \item $\dim {}^\Gamma C^{-\infty}(\Lambda,V(\sigma_\mu)) <\infty$ for all $\mu\in PS_{res}(\sigma)$. \item The singularities of $ext:C^{-\infty}(B,V_B(\sigma_\mu))\rightarrow {}^\Gamma C^{-\infty}(\partial X,V(\sigma_\mu))$ are isolated in $\{\mu\in{\aaaa_\C^\ast}\:|\:{\rm Re }(\mu)\ge 0\}$. \end{enumerate} \end{prop} {\it Proof.$\:\:\:\:$} Let $L^2(X,V(\gamma))_d$ denote the discrete subspace of $L^2(X,V(\gamma))$. \begin{lem}\label{mmm2} There exist finitely many irreducible (hence finite-dimensional) mutually inequivalent representations $(\chi_i,W_i)$ of $D_\gamma$ such that \begin{equation}\label{mmm1} L^2(X,V(\gamma))_d\cong\bigoplus_{i}V_{\pi_i}\otimes W_i \end{equation} as a $G\times D_\gamma$-module. Here $V_{\pi_i}$ are representations of the discrete series $\hat{G}_d$ of $G$. Given $\sigma\in\hat{M}$ and $\lambda\in{\aaaa_\C^\ast}$, ${\rm Re }(\lambda)>0$, let $\gamma$ be a minimal $K$-type of the principal series representation $C^\infty(\partial X,V(\sigma_\lambda))$ of $G$. Then $\chi_i\not\cong \chi_{\sigma,\lambda}$, $\forall i$. \end{lem} {\it Proof.$\:\:\:\:$} The Harish-Chandra Plancherel Theorem for $L^2(G)$ implies that $$L^2(X,V(\gamma))_d=\bigoplus_{\pi\in \hat{G}_d} V_\pi\otimes {\mbox{\rm Hom}}_K(V_\pi,V_\gamma)\ .$$ Since the representations $V_\pi$ are irreducible and mutually non-equivalent the same is true for the ${\cal U}({\bf g})^K$-modules ${\mbox{\rm Hom}}_K(V_\pi,V_\gamma)$ (see \cite{wallach88}, 3.5.4.). Since there is only a finite number of $\pi\in \hat{G}_d$ with ${\mbox{\rm Hom}}_K(V_\pi,V_\gamma)\not=0$, equation (\ref{mmm1}) follows. Let $\gamma$ now be a minimal $K$-type of $C^\infty(\partial X,V(\sigma_\lambda))$. Argueing by contradiction we assume that $\chi_{\sigma,\lambda}\cong\chi_i$ for some $i$. Then there is an embedding $$V_{\pi_i}\otimes W_i\hookrightarrow {\cal E}_{\sigma,\lambda}(X,V(\gamma))\cap L^2(X,V(\gamma))=:L^2_{\sigma,\lambda}\ .$$ There exists $f\in L^2_{\sigma,\lambda}$ such that $f(1)\not=0$. Let $f_\gamma$ be its projection onto the (left) $K$-type $\gamma$. Then $f_\gamma\not=0$. But $f_\gamma$ is the Poisson transform of some $\phi\in C^\infty(\partial X,V(\sigma_\lambda))(\gamma)$ (see \cite{olbrichdiss}, Thm. 3.6, and \cite{minemura92}). Thus $f_\gamma=P^T_\lambda(\phi)$ and asymptotically $$f_\gamma(ka)\stackrel{a\to\infty}{\sim} c_\sigma(\lambda) T\phi(k) a^{\lambda-\rho}\ .$$ Since ${\rm Re }(\lambda)>0$ we have $c_\sigma(\lambda)\not=0$. We conclude that $f_\gamma\not\in L^2$. This is a contradiction to $f_\gamma\in L^2_{\sigma,\lambda}$. \hspace*{\fill}$\Box$ \\[0.5cm]\noindent For the following two lemmas let $\gamma$ be a minimal $K$-type of $C^\infty(\partial X,V(\sigma_\lambda))$. Then its multiplicity is one. By Frobenius reciprocity $[\gamma:\sigma]=1$ and $\chi_{\sigma,\lambda}$ is a one-dimensional representation. \begin{lem}\label{mko1} Let ${\rm Re }(\lambda)>0$. If there exist $\sigma^\prime\in\hat{M}$, $\lambda^\prime\in{\aaaa_\C^\ast}$, ${\rm Re }(\lambda^\prime)\ge 0$, such that ${\cal E}_{\sigma,\lambda}(X,V(\gamma))\cap {\cal E}_{\sigma^\prime,\lambda^\prime}(X,V(\gamma))\not= 0$, then ${\rm Re }(\lambda^\prime)\ge{\rm Re }(\lambda)$. \end{lem} {\it Proof.$\:\:\:\:$} As in the proof of Lemma \ref{mmm2} there exists $$0\not=f_\gamma\in \left({\cal E}_{\sigma^\prime,\lambda^\prime}(X,V(\gamma))\cap {\cal E}_{\sigma^\prime,\lambda^\prime}(X,V(\gamma))\right)(\gamma)\ .$$ Again $f_\gamma=P^T_\lambda(\phi)=P^{T^\prime}_{\lambda^\prime}(\phi^\prime)$ for $\phi\in C^\infty(\partial X,V(\sigma_\lambda))(\gamma)$, $\phi^\prime\in C^\infty(\partial X,V({\sigma}^\prime_{\lambda^\prime}))(\gamma)$. Thus on the one hand we have $$f_\gamma(ka)\stackrel{a\to\infty}{\sim} c_\sigma(\lambda) T\phi(k) a^{\lambda-\rho}\ ,$$ and on the other hand for any $\epsilon>0$ $$f_\gamma(ka)\stackrel{a\to\infty}{\sim} o(a^{\lambda^\prime-\rho+\epsilon})\ .$$ This implies ${\rm Re }(\lambda^\prime)\ge{\rm Re }(\lambda)$.\hspace*{\fill}$\Box$ \\[0.5cm]\noindent \begin{lem}\label{o9o9o} If $\lambda\in{\bf a}^*$, $\lambda>0$, then there exists a commutative algebra extension ${\cal A}\subset{\cal D}_\gamma$ of ${\cal Z}_\gamma$ which is generated by selfadjoint elements such that the character $(\chi_{\sigma,\lambda})_{|{\cal A}}$ does not belong to the spectrum of ${\cal A}$ on $L^2(X,V(\gamma))$. \end{lem} {\it Proof.$\:\:\:\:$} Consider the finite set $B_{\sigma,\lambda}:=\{(\sigma^\prime,\lambda^\prime)\:|\: \sigma^\prime \subset \gamma_{|M}, \lambda^\prime\in{\aaaa_\C^\ast}, (\chi_{\sigma^\prime,\lambda^\prime})_{|{\cal Z}}=(\chi_{\sigma,\lambda})_{|{\cal Z}}\}$. If $(\sigma^\prime,\lambda^\prime)\in B_{\sigma,\lambda}$, then $\lambda^\prime\in{\bf a}^*$ since $(\chi_{\sigma,\lambda})_{|{\cal Z}}$ is a real character. We define $B_{\sigma,\lambda}^+:=\{(\sigma^\prime,\lambda^\prime)\in B_{\sigma,\lambda}\:|\: {\rm Re }(\lambda^\prime)=0\}$. Note that the $\chi_i$ (the $\chi_i$ have been introduced in Lemma \ref{mmm2}) and the $\chi_{\sigma^\prime,\lambda^\prime}$, $(\sigma^\prime,\lambda^\prime)\in B_{\sigma,\lambda}^+ $ are $*$-representations of ${\cal D}_\gamma$. While the $\chi_i$ were already irreducible the $\chi_{\sigma^\prime,\lambda^\prime}$, $(\sigma^\prime,\lambda^\prime)\in B_{\sigma,\lambda}^+$, can be completely decomposed into irreducible components. Let $\chi^+$ denote the representation of ${\cal D}_\gamma$, which is obtained by taking the direct sum of the $\chi_i$ and mutually inequivalent representatives of the irreducible components of $\chi_{\sigma^\prime,\lambda^\prime}$, $(\sigma^\prime,\lambda^\prime)\in B_{\sigma,\lambda}^+$. Set $I_{\sigma,\lambda}:=\ker\chi_{\sigma,\lambda}\subset {\cal D}_\gamma$ and $I^+:=\ker\chi^+\subset {\cal D}_\gamma$. We claim that $I^+\not\subset I_{\sigma,\lambda}$. Argueing by contradiction we assume that $I^+\subset I_{\sigma,\lambda}$. Let $R^+$ denote the range of the representation $\chi^+$. Since $\chi^+$ is the direct sum of mutually inequivalent representations of ${\cal D}_\gamma$ the commutant of $R^+$ is generated by the projections onto these components. Thus $R^+$ is a finite direct sum $\oplus_j {\rm Mat}(l_j,{\bf C})$ of matrix algebras. $R^+$ admits a character $\kappa:R^+\rightarrow {\bf C}$ such that the composition ${\cal D}_\gamma\rightarrow R^+\stackrel{\kappa}{\rightarrow} {\bf C}$ coincides with $\chi_{\sigma,\lambda}$. If $l_j>1$, then the restriction of $\kappa$ to the summand ${\rm Mat}(l_j)$ vanishes. Thus the representation $\chi_{\sigma,\lambda}$ of ${\cal D}_\gamma$ must be one of the one-dimensional components defining $\chi^+$. By Lemma \ref{mmm2} we have $\chi_{\sigma,\lambda}\not=\chi_i$ for all $i$. Hence there exists $(\sigma^\prime,\lambda^\prime)\in B_{\sigma,\lambda}^+$ such that $\chi_{\sigma^\prime,\lambda^\prime}$ contains $\chi_{\sigma,\lambda}$ as an irreducible component. But then ${\cal E}_{\sigma,\lambda}(X,V(\gamma))\subset {\cal E}_{\sigma^\prime,\lambda^\prime}(X,V(\gamma))$. By Lemma \ref{mko1} we conclude ${\rm Re }(\lambda^\prime)\ge \lambda>0$. This is in conflict with the definition of $B_{\sigma,\lambda}^+$. Since the ideal $I^+$ is a $*$-ideal there exists a selfadjoint $A\in {\cal D}_\gamma$ with $A\in I^+\setminus I_{\sigma,\lambda}$. Let ${\cal A}$ be the algebra generated by $A$ and ${\cal Z}_\gamma$. Let $f\in S(X,V(\gamma))$ be an eigenfunction of ${\cal A}$ corresponding to $(\chi_{\sigma,\lambda})_{|{\cal A}}$. The Harish-Chandra Plancherel theorem for the Schwartz space $S(X,V(\gamma))$ (see \cite{arthur75}) implies that $f=f_c+f_d$, where $f_d\in L^2(X,V(\gamma))_d$ and $$f_c\in\sum_{(\sigma^\prime,\lambda^\prime)\in B_{\sigma,\lambda}^+} {\cal E}_{\sigma^\prime,\lambda^\prime}(X,V(\gamma))\ .$$ It follows that $$f=\frac{1}{\chi_{\sigma,\lambda}(A)} A f = \frac{1}{\chi_{\sigma,\lambda}(A)} (Af_c + A f_d)=0\ .$$ We conclude that there are no tempered eigenfunctions of ${\cal A}$ for the character $\chi_{\sigma,\lambda}$. Thus $\chi_{\sigma,\lambda}$ is not on the spectrum of ${\cal A}$ on $L^2(X,V(\gamma))$. This finishes the proof of the lemma.\hspace*{\fill}$\Box$ \\[0.5cm]\noindent Now we finish the proof of Proposition \ref{ddsp}. Fix $\sigma\in \hat{M}$. Let $\gamma$ be a minimal $K$-type of the principal series representation associated to $\sigma$. We choose an $M$-equivariant embedding $T:V_\sigma\rightarrow V_\gamma$. If $\mu,\lambda\in{\bf a}^+$, $\lambda>\mu>0$, and if $\phi\in {}^\Gamma C^{-\infty}(\Lambda,V(\sigma_\lambda))$, $\psi\in {}^\Gamma C^{-\infty}(\Lambda,V(\sigma_\mu))$, then $P^T_\lambda(\phi)\perp P^T_\mu(\psi)$ in $L^2(Y,V_Y(\gamma))$. If ${\rm Re }(\mu)>0$, then $c_\sigma(\mu)\not=0$ and the Poisson transform $P^T_\mu$ is injective. By Corollary \ref{nahenull} there exists $\epsilon>0$ such that $PS_{res}(\sigma)\cap(0,\epsilon)=\emptyset$. In the following we argue by contradiction. Assume that $\oplus_{\lambda\in [\epsilon,\rho]} {}^\Gamma C^{-\infty}(\Lambda,V(\sigma_\lambda))$ is infinite-dimensional. Since $[\epsilon,\rho]$ is compact this would imply that there exists a sequence $\mu_i$ and $\psi_i\in {}^\Gamma C^{-\infty}(\Lambda,V(\sigma_\mu))$, such that $\mu_i\to \mu\in [\epsilon,\rho]$ and the $\psi$ are pairwise orthonormal. Let $\chi:=\chi_{\sigma,\mu}$ and let ${\cal A}$ be the algebra constructed for $\chi$ in Lemma \ref{o9o9o} such that $\chi$ does not belong to the essential spectrum of ${\cal A}$ on $L^2(X,V(\gamma))$. But our assumption implies that $\chi$ belongs to the essential spectrum of ${\cal A}$ on $L^2(Y,V_Y(\gamma))$. In fact for any $A\in{\cal A}$ we have $$\lim_{i\to\infty}\|(A-\chi(A))\psi_i\|\le \lim \lim_{i\to\infty}\|(A-\chi_{\sigma,\mu_i}(A))\psi_i\|+\lim_{i\to\infty}\|(\chi_{\sigma,\lambda}(A)-\chi(A))\psi_i\|=0\ .$$ By Proposition \ref{esspec} we conclude that $\chi$ belongs to the essential spectrum of ${\cal A}$ on $L^2(Y,V_Y(\gamma))$. But this contradicts our construction of ${\cal A}$. Thus $\oplus_{\lambda\in [\epsilon,\rho]} {}^\Gamma C^{-\infty}(\Lambda,V(\sigma_\lambda))$ is finite-dimensional. This shows $1.$ and $2.$ of Proposition \ref{ddsp}. Assertion $3.$ follows from $1.$ and Lemma \ref{lead}.\hspace*{\fill}$\Box$ \\[0.5cm]\noindent We now investigate the fine structure of the discrete subspace $L^2(Y,V_Y(\gamma))_d$ for any $\gamma\in\hat{K}$. By definition $L^2(Y,V_Y(\gamma))_d$ is the closure of the subspace $L^2(Y,V_Y(\gamma))_{\cal Z}$ of ${\cal Z}$-finite vectors. As explained in Section \ref{relsec}, if $f\in L^2(Y,V_Y(\gamma))_{\cal Z}$, then it has an asymptotic expansion at infinity. Since by Theorem \ref{contsp} the eigencharacters of ${\cal Z}$ on $L^2(Y,V_Y(\gamma))_d$ belong to $PS$ they are real. This implies (see \cite{knapp86}, Ch.8) that the set of leading exponents $E(f)$ is contained in ${\bf a}^*$. To be precise with the zero exponent we distinguish two types of leading exponents $\mu=0$ which we write as $0_0,0_1$. We say that $f$ has the leading exponent $\mu=0_1$ ($\mu=0_0$) if $p(f,0,0)$ is a non-constant (constant) polynomial on ${\bf a}$. We use the leading exponents in order to define a filtration $F_*$ of $L^2(Y,V(\gamma))_{{\cal Z}}$. For any exponent $\mu$ we set $$F_\mu L^2(Y,V(\gamma))_{{\cal Z}}:=\{f\in L^2(Y,V(\gamma))_{{\cal Z}}\:|\:\mu \ge\lambda \quad \forall \lambda \in E(f)\}\ .$$ Then $F_\mu L^2(Y,V(\gamma))_{{\cal Z}}$ is the subspace of $L^2(Y,V(\gamma))_{{\cal Z}}$ on which the boundary value map $p(.,\mu,0)$ is well-defined (if $\mu=0_1$, then we consider the leading coefficient $p_1(.,\mu,0)=:p(.,0_1,0)$ of $p(.,\mu,0)$). We define for any leading exponent $$L^2(Y,V_Y(\gamma))_{{\cal Z}}^\mu:=F_\mu L^2(Y,V(\gamma))_{{\cal Z}}\cap \ker(p(.,\mu,0))^\perp\ .$$ \begin{ddd}\label{t6r} We define \begin{eqnarray*} L^2(Y,V_Y(\gamma))_{res}&:=&\bigoplus_{\mu>0} L^2(Y,V_Y(\gamma))_{{\cal Z}}^\mu\\ L^2(Y,V_Y(\gamma))_{cusp}&:=&\bigoplus_{\mu<0} L^2(Y,V_Y(\gamma))_{{\cal Z}}^\mu\\ L^2(Y,V_Y(\gamma))_{scat}&:=&\bigoplus_{\mu=0_0,0_1} L^2(Y,V_Y(\gamma))_{{\cal Z}}^\mu\ . \end{eqnarray*} \end{ddd} We apriori have $$L^2(Y,V_Y(\gamma))_d=\overline{L^2(Y,V_Y(\gamma))_{res}\oplus L^2(Y,V_Y(\gamma))_{cusp}\oplus L^2(Y,V_Y(\gamma))_{scat}}\ ,$$ but by $1.$ of Theorem \ref{poinye} the subspaces defined above are already closed. We now describe these spaces in detail. \begin{theorem}\label{poinye} \begin{enumerate} \item The spectrum of ${\cal Z}$ on $L^2(Y,V_Y(\gamma))_d$ is finite. In particular $L^2(Y,V_Y(\gamma))_{*}$, $*\in\{res,cusp,scat\}$, are closed subspaces. \item The space $L^2(Y,V_Y(\gamma))_{res}$ is finite-dimensional. There is an embedding $$L^2(Y,V_Y(\gamma))_{res}\hookrightarrow \bigoplus_{\sigma\subset \gamma_{|M}} \bigoplus_{\mu \in PS(\sigma)} {}^\Gamma C^{-\infty}(\Lambda,V(\sigma_{\mu}))\otimes {\mbox{\rm Hom}}_M(V_\sigma,V_\gamma)\ .$$ \item The space $L^2(Y,V_Y(\gamma))_{cusp}$ is trivial or infinite-dimensional. More precisely, for any discrete series representation $\pi\in\hat{G}_d$ there exists an infinite-dimensional subspace $V_{\pi,\Gamma}\subset{}^\Gamma V_{\pi,-\infty}$ ($V_{\pi,-\infty}$ denotes the distribution vector globalization) such that for any $\gamma\in\hat{K}$ \begin{equation}\label{y7y7}L^2(Y,V_Y(\gamma))_{cusp}\cong\bigoplus_{\pi\in\hat{G}_d} V_{\pi,\Gamma}\otimes {\mbox{\rm Hom}}_K(V_\pi,V_\gamma)\ .\end{equation} \item There exists an embedding $$L^2(Y,V_Y(\gamma))_{scat}\hookrightarrow \bigoplus_{\sigma\subset \gamma_{|M}} \bigoplus_{0_0,0_1} {}^\Gamma C^{-\infty}(\Lambda,V(\sigma_0))\otimes {\mbox{\rm Hom}}_M(V_\sigma,V_\gamma)\ .$$ The space $L^2(Y,V_Y(\gamma))_{scat}$ is finite-dimensional. \end{enumerate} \end{theorem} {\it Proof.$\:\:\:\:$} The assertion $1.$ follows from $2.$ and (\ref{y7y7}). Assertion $2.$ follows from Lemma \ref{wo} and Proposition \ref{ddsp}. In fact, $\oplus_{\mu>0} p(.,\mu,0)$ defines the embedding $$L^2(Y,V_Y(\gamma))_{res}\hookrightarrow \bigoplus_{\sigma\subset \gamma_{|M}} \bigoplus_{\mu \in PS(\sigma)} {}^\Gamma C^{-\infty}(\Lambda,V(\sigma_\mu))\otimes {\mbox{\rm Hom}}_M(V_\sigma,V_\gamma)\ .$$ We now prove $3$. For $\pi\in \hat{G}_d$ set $V_{\pi,\Gamma}:={\mbox{\rm Hom}}_G(V_\pi^*,L^2(\Gamma\backslash G))$. For any Banach representation of $G$ on $V$ let $V_\infty$ denote the space of smooth vectors. We have \begin{eqnarray*} {\mbox{\rm Hom}}_G(V_\pi^*,L^2(\Gamma\backslash G))&=&{\mbox{\rm Hom}}_G((V_\pi)^*_\infty,L^2(\Gamma\backslash G)_\infty)\\ &\subset&{\mbox{\rm Hom}}_G((V_\pi)^*_\infty,C^\infty(\Gamma\backslash G))\\ &=&{\mbox{\rm Hom}}_\Gamma((V_\pi)^*_\infty,{\bf C})\\ &=&{}^\Gamma V_{\pi,-\infty}\ . \end{eqnarray*} Let $I_\pi$ be the natural injection $$I_\pi:V_{\pi,\Gamma}\otimes {\mbox{\rm Hom}}_K(V_\pi,V_\gamma)\hookrightarrow (L^2(\Gamma\backslash G)\otimes V_\gamma)^K=L^2(Y,V_Y(\gamma))\ .$$ The range of $I_\pi$ consists of eigenfunctions of ${\cal Z}$ which are matrix coefficients of the discrete series representation $V_\pi$. Matrix coefficients of discrete series representations are characterizwed by the fact that all their leading exponents are negative. We conclude that $I_\pi$ injects into $L^2(Y,V(\gamma))_{cusp}$. Conversely, if $f\in L^2(Y,V(\gamma))_{cusp}$, then all its leading exponents are negative and hence $f$ is a finite sum of matrix coefficients of discrete series representations. Given $\gamma$ the set $\{\pi_i\}$ of discrete series representations of $G$ satisfying $I_\pi\not=0$ is finite. Thus \begin{equation}\label{scd}L^2(Y,V(\gamma))_{cusp}=\bigoplus_i {\mbox{\rm im}}(I_{\pi_i})\end{equation} and the spectrum of ${\cal Z}$ on $L^2(Y,V(\gamma))_{cusp}$ is finite. This finishes the proof of $1$. To prove $3.$ it remains to show that $V_{\pi,\Gamma}$ is infinite-dimensional. Let $\gamma$ be a minimal $K$-type of the discrete series representation $V_\pi$ ( or more general a $K$-type occuring with multiplicity one in $V_\pi$). Let $\chi$ denote the corresponding character of ${\cal D}_\gamma$ induced by the representation of ${\cal U}({\bf g})^K$ on ${\mbox{\rm Hom}}_K(V_\pi,V_\gamma)$. Furthermore, let $\chi_i$ denote the irreducible representations of ${\cal D}_\gamma$ introduced in Lemma \ref{mmm2}. Without loss of generality we can assume that $\chi=\chi_1$. Let $\{\chi_1,\dots,\chi_r\}$ denote the subset of these representations satisfying $(\chi_i)_{|{\cal Z}}=\chi_{|{\cal Z}}$. We claim that there exists an abelian extension ${\cal A}\subset{\cal D}_\gamma$ of ${\cal Z}_\gamma$ which separates $\chi$ from all the characters occcuring in $(\chi_i)_{|{\cal A}}$, $\forall i>1$. Let $\chi^+$ denote the sum of all $\chi_i$ with $i>1$. Since the $\chi_i$ are all mutually inequivalent, the range of $\chi^+$ is a finite sum of matrix algebras $\oplus_{i>1}{\rm Mat}(l_i,{\bf C})$. Let $I^+\subset {\cal D}_\gamma$, $I\subset {\cal D}_\gamma$ denote the kernels of $\chi^+$, $\chi$. We claim that $I^+\not\subset I$. Assuming the contrary $R^+$ would admit an character $\kappa:R^+\rightarrow {\bf C}$ such that $\chi$ is given by $\chi:{\cal D}_\gamma\rightarrow R^+\stackrel{\kappa}{\rightarrow} {\bf C}$. But this is impossible by the definition of $\chi^+$. Choose a selfadjoint $A\in I^+\setminus I$ and let ${\cal A}={\cal Z}_\gamma[A]$. Then $\chi_i(A)=0$ for all $i>1$ but $\chi(A)\not=0$. Now $\chi_{|{\cal A}}$ belongs to the essential spectrum of ${\cal A}$ on $L^2(X,V(\gamma))$. By Proposition \ref{esspec} the character $\chi_{|{\cal A}}$ belongs to the essential spectrum of ${\cal A}$ on $L^2(Y,V_Y(\gamma))$. The characters $\chi_{|{\cal Z}}$ and $\chi_{|{\cal A}}$ are separated from the continuous spectrum on $L^2(X,V(\gamma))$. By Theorem \ref{contsp} the character $\chi_{|{\cal A}}$ is also separated from the continuous spectrum of ${\cal A}$ on $L^2(Y,V_Y(\gamma))_c$ and from the spectrum on $L^2(Y,V_Y(\gamma))_{scat}$. Since the discrete spectrum of ${\cal Z}$ and of ${\cal A}$ on $L^2(Y,V_Y(\gamma))$ is finite the eigenspace of ${\cal A}$ in $L^2(Y,V_Y(\gamma))$ according to $\chi_{|{\cal A}}$ must be infinite-dimensional. Since $L^2(Y,V_Y(\gamma))_{res}$ can only contribute an finite-dimensional subspace to this eigenspace the eigenspace of ${\cal A}$ to $\chi_{|{\cal A}}$ in $L^2(Y,V_Y(\gamma))_{cusp}$ is infinite-dimensional. By (\ref{scd}) this eigenspace is just given by $I_\pi (V_{\pi,\Gamma}\otimes {\mbox{\rm Hom}}_K(V_\pi,V_\gamma))$. It follows that $\dim\:V_{\pi,\Gamma}=\infty$. This finishes the proof of $3.$. We now prove $4$. There is an embedding $$p(.,0_0,0)\oplus p(.,0_1,0):L^2(Y,V_Y(\gamma))_{scat}\hookrightarrow {}^\Gamma C^{-\infty}(\partial X,V(\gamma_{|M,0}))\oplus {}^\Gamma C^{-\infty}(\partial X,V(\gamma_{|M,0}))\ .$$ We prove the assertion about the support. We show that if $f\in L^2(Y,V_Y(\gamma))_{{\cal Z}}^{0_1}$, then ${\mbox{\rm supp}}(p(f,0_1,0))\subset \Lambda$. The argument is similar to the one used in the proof of Lemma \ref{wo}. Let $U\subset \bar{U}\subset \Omega$ be open. If $\phi\in C_c^\infty(U,V(\gamma_{|M,0}))$, we have $$(\phi,p(f,0_1,0))=\lim_{a\to\infty} a^{\rho-\mu}|\log(a)|^{-1} \int_K (\phi(k),f(ka)) dk\ .$$ Constructing the sequence $\phi_n$ with ${\mbox{\rm supp}}(\phi_n)\subset UMA^+K$ as in the proof of Lemma \ref{wo} but using $\chi_n(a):=|\log(a)|^{-1}a^{-\rho-\bar{\mu}}\chi(|\log(a)|-n)$ we can write $$(\phi,p(f,0_1,0))=\lim_{n\to\infty} (\phi_n,f)\ .$$ By construction $\phi_n\to 0$ weakly in $L^2(UMA^+K,V(\gamma))$ and $f_{|UMA^+K}\in L^2(UMA^+K,V(\gamma))$. We obtain $(\phi,p(f,0_1,0))=0$. Since $U$ and $\phi$ were arbitrary, this proves that ${\mbox{\rm supp}}(p(f,0_1,0))\subset \Lambda$. The same argument works in the case $\mu=0_0$. The following lemma (for $\lambda=0$) implies that $L^2(Y,V_Y(\gamma))_{scat}$ is finite-dimensional. For $\lambda>0$ it provides an alternative proof of Proposition \ref{ddsp}, $2$. \begin{lem} If $\lambda\ge 0$, then ${}^\Gamma C^{-\infty}(\Lambda,V)(\sigma_\lambda))$ is finite-dimensional. \end{lem} {\it Proof.$\:\:\:\:$} We first prove an apriori estimate of the order of elements of ${}^\Gamma C^{-\infty}(\Lambda,V(\sigma_\lambda))$, i.e. we show that there exists a $k\in{\bf N}$ such that $f\in {}^\Gamma C^{-\infty}(\Lambda,V(\sigma_\lambda))$ implies $f\in C^k(\partial X,V(\tilde{\sigma}_{-\lambda}))^\prime$. The inclusion ${}^\Gamma C^{-\infty}(\Lambda,V(\sigma_\lambda))\hookrightarrow C^k(\partial X,V(\tilde{\sigma}_{-\lambda}))^\prime$ induces on ${}^\Gamma C^{-\infty}(\Lambda,V(\sigma_\lambda))$ the structure of a Banach space. Since ${}^\Gamma C^{-\infty}(\Lambda,V(\sigma_\lambda))$ is a closed subspace of the Montel space $C^{-\infty}(\partial X,V(\sigma_\lambda))$ it must be finite-dimensional. It remains to prove the apriori estimate of the order. If $\lambda\not=0$ or if $\lambda=0$ and the principal series representation $\pi^{\sigma,\lambda}$ is irreducible, then let $\gamma$ be a minimal $K$-type of the principal series representation $\pi^{\sigma,\lambda}$ and fix an unitary $T\in {\mbox{\rm Hom}}_M(V_\sigma,V_\gamma)$. If $\lambda=0$ and $\sigma=\sigma^\prime\oplus\sigma^{\prime w}$ is reducible, then let $\gamma$ be the sum of two copies of a minimal $K$-type of the principal series representation $\pi^{0,\sigma^\prime}$. In this case we let $T:=T^\prime_0\oplus T^\prime_1$, where $T^\prime_0\in {\mbox{\rm Hom}}_M(V_{\sigma^\prime},V_\gamma)$ and $T^\prime_1\in {\mbox{\rm Hom}}_M(V_{\sigma^{w \prime}},V_\gamma)$ are unitary. In the remaining case where $\lambda=0$, $\sigma$ is irreducible and $\pi^{\sigma,\lambda}$ splits as a sum of two irreducible representations we let $\gamma$ be the sum of minimal $K$-types of these representations. In this case we let $T\in {\mbox{\rm Hom}}_M(V_\sigma,V_\gamma)$ denote the "diagonal" embedding of $\sigma$ into $\gamma$. In any case let $P:=P^T_\lambda$ denote the associated injective Poisson transform. Consider $f\in {}^\Gamma C^{-\infty}(\Lambda,V(\sigma_\lambda))$. The asymptotic expansion (\ref{epan}) shows that $Pf$ is bounded along $\Omega$. Using the $\Gamma$-invariance we conclude that $Pf$ is a uniformly bounded section of $V(\gamma)$. Let $\chi$ be the infinitesimal character of ${\cal Z}$ on the principal series representation $\pi^{\sigma,\lambda}$. Let $C^\infty_{mg}(X,V(\gamma))_\chi$ denote the corresponding eigenspace. As a topological vector space $C^\infty_{mg}(X,V(\gamma))_\chi$ is a direct limit of Banach spaces $$C^\infty_{R}(X,V(\gamma))_\chi:=\{f\in C^\infty_{mg}(X,V(\gamma))_\chi\:|\: \sup_{g\in G}\|g\|^{-R} |f(g)| <\infty\}\ .$$ In particular, $Pf\in C^\infty_{0}(X,V(\gamma))_\chi$. The range of the Poisson transform $P$ is a closed $G$-submodule ${\cal M}$ of $C^\infty_{mg}(X,V(\gamma))_\chi$. We claim that there is a boundary value map $\beta$ defined on ${\cal M}$ which is continuous and inverts $P$. Before proving the claim we finish the proof of the apriori estimate assuming the claim. On the one hand the topological vector space ${\cal M}$ is the direct limit of the Banach spaces ${\cal M}_R$, where ${\cal M}_R:={\cal M}\cap C^\infty_{R}(X,V(\gamma))_\chi$. On the other hand $C^{-\infty}(\partial X, V(\sigma_\lambda))$ is the direct limit of Banach spaces $C^k(\partial X,V(\tilde{\sigma}_{-\lambda}))^\prime$. Since $\beta$ is continuous for any $R\ge 0$ there exists a $k\in{\bf N}$ such that $\beta({\cal M}_R)\subset C^k(\partial X,V(\tilde{\sigma}_{-\lambda}))^\prime$. Since $Pf\in {\cal M}_0$ this yields the apriori estimate we looked for. We now show the existence of the boundary value $\beta$. It is intimately related with the leading asymptotic coefficient $p(\phi,\lambda,0)$, $\phi\in {\cal M}$. Let first $\lambda>0$. Then we can define $\beta(\phi)$, $\phi\in {\cal M}$, by $$( \beta(\phi),\psi) := c_\sigma(\lambda)^{-1}\lim_{a\to\infty} a^{-\lambda+\rho} \int_K (\phi(ka),T\psi(k)) dk, \quad \psi\in C^\infty(\partial X,V(\sigma_{-\lambda}))\ .$$ Let now $\lambda=0$. If $c_\sigma(\mu)$ has a pole at $\mu=0$, then $\pi^{\sigma,\lambda}$ is irreducible an we define $\beta(\phi)$, $\phi\in {\cal M}$, by $$(\beta(\phi),\psi) := \frac{1}{2{\rm res}_{\mu=0}c_\sigma(\mu)} \lim_{a\to\infty}\log(a)^{-1} a^{\rho} \int_K (\phi(ka),T\psi(k) ) dk, \quad \psi\in C^\infty(\partial X,V(\sigma_{0})) .$$ If $c_\sigma(\mu)$ is regular at $\mu=0$ and $\sigma=\sigma^\prime\oplus\sigma^{\prime w}$, then we define $\beta$ by $$(\beta(\phi),\psi) := \frac{1}{c_\sigma(0)} \lim_{a\to\infty} a^{\rho} \int_K (\phi(ka),T\psi(k) ) dk, \quad \psi\in C^\infty(\partial X,V(\sigma_{0})) .$$ In the remaining case $c_\sigma(\mu)$ is regular at $\mu=0$ and $\pi^{\sigma,0}$ is reducible. Let $\gamma=\gamma_1\oplus\gamma_2$ and $t_i\in{\mbox{\rm Hom}}_M(V_\sigma,V_{\gamma_i})$, $i=1,2$, be such that $T=t_1\oplus t_2$. Let $c_{\gamma_i,\sigma}(\mu)$ denote the value of $c_{\gamma_i}(\mu)$ on the range of $t_i$. Note that $c_{\gamma_i,\sigma}(0)\not=0$. We define $\beta$ by $$(\beta(\phi),\psi) := \lim_{a\to\infty} a^{\rho} \int_K (\phi(ka),(t_1 c_{\gamma_1,\sigma}(0)^{-1}\oplus t_2 c_{\gamma_2,\sigma}(0)^{-1}) \psi(k) ) dk, \quad \psi\in C^\infty(\partial X,V(\sigma_{0})) .$$ One can check in each case that $\beta$ inverts the Poisson transform $P$ (e.g. use the method of the proof of \cite{olbrichdiss}, Lemma 4.31). The fact that $\beta$ is continuous follows from the globalization theory for Harish-Chandra modules. \hspace*{\fill}$\Box$ \\[0.5cm]\noindent This finishes the proof of $4.$ and hence of the theorem. \hspace*{\fill}$\Box$ \\[0.5cm]\noindent It is clear by Theorem \ref{contsp} that $L^2(Y,V_Y(\gamma))_c$ is always non-trivial. By Theorem \ref{poinye} there are examples with $L^2(Y,V_Y(\gamma))_{cusp}$ non-trivial. If $\delta_\Gamma>0$, then the Patterson-Sullivan measure leads to a non-trivial element in $L^2(Y)_{res}$. Thus examples with non-trivial $L^2(Y,V_Y(\gamma))_{res}$ exist. We now give an example with $L^2(Y,V_Y(\gamma))_{scat}\not=0$. Let $\Gamma\subset SO(1,2)$ be a cocompact Fuchsian group. Consider $\Gamma\subset SO(1,3)$ in the standard way. Then $Y$ is a $3$-dimensional hyperbolic manifold of the type considered in the present paper. $Y$ has two ends, i.e., $B$ has two components. It was shown by Mazzeo-Phillips \cite{mazzeophillips90}, Corollary 3.20, that the dimension of the space of harmonic, square-integrable one-forms is $\ge \sharp\{\mbox{ends of $Y$}\}-1=1$. The character of ${\cal Z}$ corresponding to harmonic one-forms is the boundary of the continuous spectrum of ${\cal Z}$ on one-forms. Thus square-integrable harmonic one-forms are elements of $L^2(Y,V_Y(\gamma))_{scat}$. As indicated after the proof of Proposition \ref{upperbound} one can obtain vanishing results for the discrete spectrum. One can bound the residual spectrum in terms of $\delta_\Gamma$. For certain $K$-types (e.g. for the trivial one) one can show that $L^2(Y,V_Y(\gamma))_{scat}$ vanishes (see the remark following the proof of Lemma \ref{th43}). \bibliographystyle{plain}
proofpile-arXiv_065-434
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction}\setcounter{equation}{0} Systems of two-dimensional electrons in a perpendicular strong magnetic field are giving many exciting physics these days\cite{r1}. Electrons in these systems have discrete energies with finite maximum degeneracy\cite{r2}. Among infinitely many representations, an invariant representation under two dimensional translation is convenient for many purposes. Invariant representation under continuous translations does not exist due to the magnetic field but lattice translational invariant one does exist. von Neumann lattice coherent state representation\cite{r3}is such representation and we have used it to verify the integer quantum Hall effect. We give a new representation of von Neumann lattice in this paper, and propose a new mean field theory of the quantum Hall liquid based on flux condensed state. The fractional Hall state\cite{r4} is regarded as a kind of integer Hall effect due to the condensed flux, and has a large energy gap in our theory. It is convenient to use relative coordinates\cite{r5} which are perpendicular to velocity operators and guiding center coordinates for describing two-dimensional electrons in the perpendicular magnetic field. Coherent states in guiding center coordinates have minimum spatial extensions and its suitable complete subset whose element has discrete eigenvalues defined on lattice site are used. Field operators in the present representation have two-dimensional lattice coordinates in addition to Landau level index, and the many-body theory is described by a lattice field theory of having internal freedom. Exact identities such as current conservation and Ward-Takahashi identity\cite{r3} derived from current conservation are written in a transparent way and play the important roles for establishing an exact low-energy theorem of the quantum Hall effect. We are able to expess flux state in a symmetric way, also. One of the hardest and most important thing of the fractional Hall effect is to find a mechanism of generating energy gap. Without interactions there is no energy gap at any fractional filling state. Hence interaction generates gap. Once the gap is generated, fluctuations are suppressed. Higher order corrections are expected to be small if starting approximate ground state has the energy gap and close to eigenstate. Flux state has a modified symmetry under translation and can have energy gap at certain filling states. We study flux states of the fractional Hall system. In our mean field theory of the fractional quantum Hall effect, an order parameter is a magnitude of the flux condensation. Because the system changes its property drastically with the change of flux it is quite natural to treat the flux as an order parameter. Due to the dynamical fux, our mean field Hamiltonian is close to that of Hofstadter\cite{r6} and new integer Hall effect occurs within the lowest Landau level space of the external magnetic field. $ $From Hofstadter's analysis and others\cite{r7} ground state has the lowest energy and the largest energy gap, if its flux is proportional to filling factor. We regard, in fact, these states as the fractional Hall states. Especially the principal series at $\nu=p/(2p\pm1)$ satisfy the self-consistency condition and have large energy gap. They can be observed even in the Hofstadter's spectrum of butterfly shape. Our mean field theory has similarities also with a mean field theory of composite fermion\cite{r8} in regarding the fractional Hall effect as a kind of integer Hall effect. However, ours includes interaction effects in the space of the lowest Landau level at the mean field level and so the effective mass, energy gap, and other quantities of the fractional quantum Hall states are close to the observed value even in the lowest order. These features are seen in ours but are not seen in the composite fermion mean field theory. We, also, found our ground state energy has slightly higher energy than that of Laughlin for $\nu=1/3$ case. Due to the energy gap, fluctuations are small in the fractional quantum Hall state. Higher order corrections are small and the perturbative expansions converge well, generally. Exact half-filling is, however, exceptional and there is no energy gap. Fermi surface is composed of isolated points in the lowest order, but the fluctuations are extremely large. So, the structure around the Fermi energy may be changed drastically from that of the lowest order by interactions at the half-filling. Thus our mean field may not be good at the exact half-filling. The paper is organized in the following way. In section 2, we formulate von Neumann lattice representation and verify the integer Hall effect. In section 3 flux state mean field theory on the von Neumann lattice is formulated and is compared with the existing experiments in Section 4. Conclusions are given in section 5. \section{Quantum Hall dynamics on von Neumann lattice}\setcounter{equation}{0} Quantum Hall system is described by the following Hamiltonian, \begin{eqnarray} &H=\int d{\bf x}[\Psi^\dagger(x){({\bf p}+e{\bf A})^2\over 2m}\Psi(x)+ \rho(x){V(x-y)\over2}\rho(y)], \\ &\partial_1 A_2-\partial_2 A_1=B,\ \rho(x)=\Psi^\dagger(x)\Psi(x), \ V(x)={e^2\over\kappa}{1\over\vert x\vert}. \nonumber \end{eqnarray} We ignore the disorders in this paper. It is convenient to use the following two sets of variables, \begin{eqnarray} &\xi={1\over eB}(p_y+eA_y),\ X=x-\xi,\\ &\eta=-{1\over eB}(p_x+eA_x),\ Y=y-\eta,\nonumber \end{eqnarray} which satisfy, \begin{eqnarray} &[\xi,\eta]=-[X,Y]=-i{\hbar\over eB},\nonumber \\ &[\xi,X]=[\xi,Y]=0,\\ &[\eta,X]=[\eta,Y]=0.\nonumber \end{eqnarray} We expand the electron field operators with a complete set of base functions, $f_l(\xi,\eta)\otimes\vert R_{m,n}\rangle$, which are defined by, \begin{eqnarray} &{({\bf p}+{\bf A})^2\over 2m}f_l={e^2 B^2\over 2m}(\xi^2+\eta^2)f_l= E_l f_l,\ E_l={\hbar eB\over 2m}(2l+1)\\ &(X+iY)\vert R_{m,n}\rangle=a(m+in)\vert R_{m,n}\rangle, \ a=\sqrt{2\pi\hbar\over eB}. \nonumber \end{eqnarray} The coherent states on von Neumann lattice are constructed as, \begin{eqnarray} &\vert R_{m,n}\rangle=(-1)^{mn+m+n}e^{A^\dagger\sqrt{\pi}(m+in)- A\sqrt{\pi}(m-in)}\vert 0\rangle,\\ &A=\sqrt{eB\over 2\hbar}(X+iY),\ [A,A^\dagger]=1. \nonumber \end{eqnarray} The expressions of electron field are given by, \begin{eqnarray} &\Psi(x)=\sum a_l(m,n)f_l(\xi,\eta)\otimes\vert R_{m,n}\rangle,\\ &\Psi^\dagger(x)=\sum a^\dagger_l(m,n) f_l(\xi,\eta)\otimes\langle R_{m,n}\vert, \nonumber \end{eqnarray} and they are substituted into the action integral, \begin{eqnarray} S&=&\int d{\bf x}\Psi^\dagger(x)i\hbar{\partial\over\partial t}\Psi(x) -H \\ &=&\sum\langle R_{m_1,n_1}\vert R_{m_2,n_2}\rangle a^\dagger_l(R_1)( i\hbar{\partial\over\partial t}-E_l)a_l(R_2)+ \int d{\bf k}\rho({\bf k}){V({\bf k})\over 2}\rho(-{\bf k}). \nonumber \end{eqnarray} By using the conjugate momentum to the lattice site, ${\bf R}_{m,n}$, the action is written as, \begin{eqnarray} S=&\sum\sum a^\dagger_l({\bf p}_1)e^{i\phi({\bf p},{\bf N})} \beta^*({\bf p}_1)\beta({\bf p}_2)( i\hbar{\partial\over\partial t}-E_l)a_l({\bf p}_2) \delta({\bf p}_1-{\bf p}_2+{2\pi\over a}{\bf N}) \nonumber\\ &+\int d{\bf k}\rho({\bf k}){V({\bf k})\over 2}\rho(-{\bf k}), \end{eqnarray} \begin{eqnarray*} \beta^*({\bf p})&=&2^{1\over4}e^{-{a^2\over 4\pi}p^2_x} \Theta_1({ia\over 2\pi}(p_x+ip_y),i),\nonumber\\ \beta^*({\bf p}+{2\pi\over a}{\bf N})&=&e^{i\phi({\bf p},{\bf N})} \beta^*({\bf p}),\nonumber\\ \phi({\bf p},{\bf N})&=&N_x(ap_y+\pi)+N_y\pi,\nonumber\\ &&-\pi/a\leq p_i\leq\pi/a, \nonumber \end{eqnarray*} where $\Theta_1(x,i)$ is the elliptic theta function of the first kind\cite{r9} and ${\bf N}$ is a two-dimensional vector which has integers as components. The momentum is conserved and its eigenvectors are orthogonal if their eigenvalues are different but they are not normalized. We introduce normalized operators, \begin{eqnarray} b_l({\bf p})&=&\beta({\bf p})a_l({\bf p}),\\ b^\dagger_l({\bf p})&=&\beta^*({\bf p})a^\dagger_l({\bf p}). \nonumber \end{eqnarray} $ $From the definition, $\beta({\bf p})$ vanishes at ${\bf p}=0$. This reflects a constraint of the coherent states on von Neumann lattice\cite{r10}, but it causes no difficulty in our method. They satisfy, \begin{eqnarray} &\{b_{l_1}({\bf p}_1),b^\dagger_{l_2}({\bf p}_2)\}= \sum_{{\bf N}}e^{-i\phi({\bf p}_1,{\bf N})}\delta({\bf p}_1-{\bf p}_2 +{2\pi\over a}{\bf N}),\\ &b_l({\bf p}+{2\pi\over a}{\bf N})=e^{-i\phi({\bf p}_1,{\bf N})}b_l( {\bf p}).\nonumber \end{eqnarray} The action is written as, \begin{eqnarray} S=\sum_l\sum_{{\bf N},{\bf p}_1,{\bf p}_2} &b^\dagger_l({\bf p}_1)e^{i\phi({\bf p}_1,{\bf N})}( i\hbar{\partial\over\partial t}-E_l)b_l({\bf p}_2) \delta({\bf p}_1-{\bf p}_2+{2\pi\over a}{\bf N}) \nonumber\\ &+\int d{\bf k}\rho({\bf k}){V({\bf k})\over 2}\rho(-{\bf k}), \end{eqnarray} \begin{equation} \rho({\bf k})=\sum b^\dagger_{l_1}({\bf p}_1)b_{l_2}({\bf p}_2) \delta({\bf p}_1-{\bf p}_2-{\bf k}+{2\pi\over a}{\bf N}) (l_1\vert e^{i{\bf k}\cdot\hbox{$\xi$\kern-.5em\hbox{$\xi$}}}\vert l_2)e^{i\phi({{\bf p}_1+ {\bf p}_2\over2},{\bf N})+i{a^2\over 4\pi}k^y(p_1^x+p_2^x)}. \end{equation} The current operator is also written as, \begin{eqnarray} &j_i({\bf k})=\sum b^\dagger_{l_1}({\bf p}_1)b_{l_2}({\bf p}_2)\delta( {\bf p}_1-{\bf p}_2-{\bf k}+{2\pi\over a}{\bf N}) (l_1\vert v_i e^{i{\bf k}\cdot\hbox{$\xi$\kern-.5em\hbox{$\xi$}}}\vert l_2)e^{i\phi({{\bf p}_1+ {\bf p}_2\over2},{\bf N}) +i{a^2\over 4\pi}k^y(p_1^x+p_2^x)}, \nonumber\\ &{\bf v}={eB\over m}(-\eta,\xi). \end{eqnarray} The commutation relations (2.10), and the action (2.11), have no singularity in ${\bf p}$. The zero of $\beta({\bf p})$ in Eq.(2.9) does not cause any problem. Hence meaningful theory is defined in this way contrary to naive expectations. To make sure this point further, we solve the one impurity problem by the present representation. We find an agreement with previous results. Namely eigenvectors are localized around the impurity position if corresponding eigenvalues are isolated and located between Landau levels. Their energies are moved with the impurity, also. Appendix A is devoted for the short range impurity problem. $ $From Eqs.(2.10) and (2.12), we have commutation relations between charge density and field operators, \begin{eqnarray} &[\rho({\bf k}),b({\bf p})]=-(l_1\vert e^{i{\bf k}\cdot\hbox{$\xi$\kern-.5em\hbox{$\xi$}}} \vert l_2) e^{i{a^2\over 4\pi}k_y(2p_x-k_x)}b_{l_2}({\bf p}-{\bf k}) e^{i\pi N_x N_y+iak_y N_x},\nonumber\\ &\vert {\bf p}-{\bf k}+{2\pi\over a}{\bf N}\vert\leq {\pi\over a},\\ &[\rho({\bf k}),b^\dagger({\bf p})]= (l_1\vert e^{i{\bf k}\cdot\hbox{$\xi$\kern-.5em\hbox{$\xi$}}}\vert l_2) e^{i{a^2\over 4\pi}k_y(2p_x+k_x)}b^\dagger_{l_1}({\bf p}+{\bf k}) e^{-i\pi N_x N_y-iak_y N_x},\nonumber\\ &\vert {\bf p}+{\bf k}-{2\pi\over a}{\bf N} \vert\leq {\pi\over a}.\nonumber \end{eqnarray} For the momentum $\bf p$ in the fundamental region and infinitesimal $\bf k$, $\bf N$ vanishes. The right-hand sides have linear terms in $\bf k$. Hence Ward-Takahashi identity between the vertex part and the propagator is modified from that of the naive one. We introduce the unitary operator $U({\bf p})$ which satisfies \begin{equation} \delta_{l_1,l_2}{\partial\over\partial p_i}U({\bf p})+ \{i(l_1\vert\xi_i\vert l_2)+i{a^2\over 2\pi}p_x\delta_{l_1,l_2}\}U(p)=0, \end{equation} and make transformation of the propagator and the vertex part as, \begin{eqnarray} &\tilde S({\bf p})=U({\bf p})S^{(0)}({\bf p})U^\dagger({\bf p}),\\ &\tilde \Gamma_\mu({\bf p}_1,{\bf p}_2)=U({\bf p}_1) \Gamma_\mu({\bf p}_1,{\bf p}_2)U^\dagger({\bf p}_2).\nonumber \end{eqnarray} They satisfy, then, \begin{equation} \tilde \Gamma_\mu(p,p)={\partial\tilde S^{-1}(p)\over\partial p_\mu}. \end{equation} The current correlation function in the momentum representation is written as, \begin{eqnarray} \pi_{\mu\nu}(q)&=&\langle{\rm T}(j_\mu(q_1)j_\nu(q_2))\rangle \\ &=&\sum S^{(0)}_{l_1,l_4}(p_1)S^{(0)}_{l_3,l_2}(p_3) (l_1\vert\Gamma_\mu e^{i{\bf q}\cdot\hbox{$\xi$\kern-.5em\hbox{$\xi$}}}\vert l_2) (l_3\vert\Gamma_\nu e^{-i{\bf q}\cdot\hbox{$\xi$\kern-.5em\hbox{$\xi$}}}\vert l_4) \times \nonumber \\ &&\delta({\bf p}_1-{\bf p}_3-{\bf p}_1+{2\pi\over a}{\bf N}) \delta(-{\bf q}_1-{\bf p}_2+{2\pi\over a}({\bf M}+{\bf N})), \nonumber\\ S^{(0)}_{l_1,l_4}(p_1)&=&{1\over p_1^0-E_{l_1}}\delta_{l_1,l_4}, \nonumber \end{eqnarray} in the lowest order of the interaction. The Hall conductance is the slope of $\pi_{\mu\nu}(q_1)$ at the origin and is expressed, under the use of the transformed propagator and the Ward-Takahashi identity (2.17), as, \begin{equation} \sigma_{xy}={e^2\over h}{1\over 24\pi^2}\int d^3 p\epsilon_{\mu\nu\rho} {\rm Tr}[{\partial\tilde S^{-1}(p)\over\partial p_\mu}\tilde S(p) {\partial\tilde S^{-1}(p)\over\partial p_\nu}\tilde S(p) {\partial\tilde S^{-1}(p)\over\partial p_\rho}\tilde S(p)]. \end{equation} The right-hand side is a three-dimensional winding number of the mapping defined by the propagator $\tilde S(p)$. The space of the propagator is decomposed into SU(2) subspace, a space spanned by $l$-th Landau level and $(l+1)$-th Landau level from Eq.(2.18). Note also that $S^{(0)}(p)$ is diagonal in $l_i$. In this subspace, $\tilde S(p)$ is defined on the torus and the Hall conductance becomes integer multiple of $e^2\over h$ in the quantum Hall regime where there is no two-dimensionally extended states around the Fermi energy. The value is stable under perturbation effects such as interactions, disorders, and others as far as the series converge, as were shown by Coleman, Hill\cite{r11}, and others\cite{r12}\cite{r3}. The value of $\sigma_{xy}$ stays constant while the Fermi energy is moved if there is no singularity involved. This occurs actually at the quantum Hall regime where there is no two-dimensionally extended states but there are only localized states with discrete energies or one-dimensionally extended states. In this situation, the value is computed by the lowest order calculation and has no correction from the higher order corrections. \section{Flux state mean field theory}\setcounter{equation}{0} We propose a new mean field theory based on flux state on von Neumann lattice in this section. The dynamical flux is generated by interactions and plays the important role in our mean field theory. It is described by a lattice Hamiltonian, which is due to the external magnetic field, and by the dynamical magnetic flux due to interaction, although the original electrons are defined on the continuum space. Consequently, our mean field Hamiltonian is close to Hofstadter Hamiltonian and hence there are similarities between their solutions. Hofstadter Hamiltonian shows remarkable structures. As is seen in Fig.1 the largest gap exists along a line $\Phi=\nu\Phi_0$ with a unit of flux $\Phi_0$. Ground state energy becomes minimum also with this flux. These facts may suggest that Hofstadter problem has some connection with the fractional Hall effect. We pursue a mean field theory of the condensed flux states in the quantum Hall system and point out that the Hofstadter problem is actually connected with the fractional Hall effect. $ $From the dynamical magnetic field, are defined new Landau levels. If integer number of these Landau levels are filled completely, the integer quantum Hall effect occurs. The ground state has a large energy gap and is stable against perturbations, just like ordinary integer quantum Hall effect. We study these states and will identify them as fractional quantum Hall states. We postulate, in the quantum Hall system of the filling factor $\nu$, the dynamical flux per plaquette and dynamical magnetic field of the following magnitudes, \begin{eqnarray} &\Phi_{\rm ind}=\nu\Phi_0,\ \Phi_0=\Phi_{\rm external\ flux},\\ &B_{\rm ind}=\nu B_0,\ B_0=B_{\rm external\ magnetic\ field}, \nonumber \end{eqnarray} where $\nu$ is the filling factor measured with the external magnetic field. We obtain a self-consistent solution with this flux. Then the integer quantum Hall effect due to induced magnetic field could occur just at filling factor $\nu$, because the density satisfies the integer Hall effect condition, \begin{eqnarray} &{eB_{\rm ind}\over 2\pi}N={eB_0\over 2\pi}\nu,\\ &N=1.\nonumber \end{eqnarray} The ground state has a large energy gap, generally. At the half-filling $\nu=1/2$, half-flux $\Phi_0/2$ is induced. This situation has been studied in detail by Lieb\cite{r13} and others\cite{r14} in connection with Hubbard model or t-J model. Lieb gave a quite general proof that the energy is optimal with half-flux at the half-filling case. We study first the state of $\nu=1/2$ and the states of $\nu=p/(2p\pm1)$, next. At $\Phi=\Phi_0/2$, band structure is that of massless Dirac field and has doubling symmetry. When even number of Landau levels of the effective magnetic field, $B_{\rm ind}-B_0/2$, are filled, ground states have large energy gaps. This occurs if the condition of the density, \begin{eqnarray} {e\over 2\pi}\vert\nu-{1\over2}\vert B_0\cdot 2p={eB_0\over2\pi}\nu,\\ \nu={p\over 2p\pm1}\ ;\ p,\ {\rm integer},\nonumber \end{eqnarray} is satisfied. A factor 2 in the left-hand side is due to doubling of states and will be discussed later. We study these states in detail based on von Neumann lattice representation. Action, Eq.(2.11), and density operator, Eq.(2.12), show that there is an effective magnetic field in the momentum space. Area of the momentum space is given by a finite value, $(2\pi/a)^2$, and the total flux is hence finite. The total flux in the momentum space is in fact unit flux. In the thermodynamic limit, in which the density in space is finite, the density in momentum space is infinite. Consequently, it is possible to make this phase factor disappear by a singular gauge transformation in the momentum space with infinitesimally small coupling. We make a singular gauge transformation of the field in the momentum space, \begin{eqnarray} &c_l({\bf p})=e^{i\tilde e\lambda({\bf p})}b_l({\bf p}), \nonumber\\ &\lambda({\bf p})={1\over 2\pi}\int\theta({\bf p}-{\bf p}')\rho({\bf p}') d{\bf p}',\\ &\tan\theta={(p-p')_y\over(p-p')_x},\nonumber \end{eqnarray} where $\rho({\bf p})$ is the density operator in the momentum space and $\tilde e$ is determined from Eqs.(B.3) and (B.4) in Appendix B. With the transformed field, the commutation relation and the charge density are expressed as, \begin{eqnarray} &\{c_{l_1}({\bf p}_1),c^\dagger_{l_2}({\bf p}_2)\}= \sum\delta_{l_1,l_2}\delta({\bf p}_1-{\bf p}_2+{2\pi\over a}{\bf N}),\\ &\rho({\bf k})=\sum c^\dagger_{l_1}({\bf p}_1)c_{l_2} ({\bf p}_2)\delta({\bf p}_1-{\bf p}_2-{\bf k}+{2\pi\over a}{\bf N}) (l_1\vert e^{i{\bf k}\cdot\hbox{$\xi$\kern-.5em\hbox{$\xi$}}}\vert l_2).\nonumber \end{eqnarray} By a Chern-Simons gauge theory in the momentum space, the gauge transformation, Eq.(3.4), is realized as is expressed in Appendix B. Here, the coupling constant $\tilde e$ is infinitesimally small, hence fluctuations of the Chern-Simons gauge field have small effect and we ignore the fluctuations. The action in the coordinate representation is given by, \begin{eqnarray} &S=\sum c^\dagger_l({\bf R})(i\hbar{\partial\over\partial t}-E_l) c_l({\bf R})-{1\over 2}\sum v_{l_1,l_2;l_3,l_4}({\bf R}_2-{\bf R}_1) c^\dagger_{l_1}({\bf R}_1)c^\dagger_{l_2}({\bf R}_1) c_{l_3}({\bf R}_2)c_{l_4}({\bf R}_2),\nonumber\\ &v_{l_1,l_2;l_3,l_4}({\bf R}_2-{\bf R}_1)=\int_{{\bf k}\neq0} d{\bf k} V({\bf k}) (l_1\vert e^{i{\bf k}\cdot\hbox{$\xi$\kern-.5em\hbox{$\xi$}}}\vert l_2) (l_3\vert e^{-i{\bf k}\cdot\hbox{$\xi$\kern-.5em\hbox{$\xi$}}}\vert l_4) e^{i{\bf k}\cdot({\bf R}_2-{\bf R}_1)}. \end{eqnarray} Hence the Hamiltonian in the lowest Landau level space is given by, \begin{eqnarray} &H=-{1\over 2}\sum v({\bf R}_2-{\bf R}_1) c_0^\dagger({\bf R}_1) c_0({\bf R}_2) c_0^\dagger({\bf R}_2) c_0({\bf R}_1), \\ &v({\bf R})= {\pi\over a}e^{-{\pi\over2}{\bf R}^2}I_0({\pi\over2}{\bf R}^2), \nonumber \end{eqnarray} where $I_0$ is zero-th order modified Bessel function. We study a mean field solution of this Hamiltonian. We have an expectation value and a mean field Hamiltonian, \begin{eqnarray} &\langle c_0^\dagger({\bf R}_1) c_0({\bf R}_2)\rangle= U_0({\bf R}_1-{\bf R}_2)e^{i\int_{{\bf R}_2}^{{\bf R}_1}{\bf A}_{\rm ind} \cdot d{\bf x}},\\ &H_{\rm mean\ field}=-\sum v({\bf R}_2-{\bf R}_1)U_0({\bf R}_1-{\bf R}_2) e^{i\int_{{\bf R}_2}^{{\bf R}_1}{\bf A}_{\rm ind}\cdot d{\bf x}} c_0^\dagger({\bf R}_2)c_0({\bf R}_1), \nonumber \end{eqnarray} and solve the equations self-consistently. The mean field Hamiltonian coincides to that of Hofstadter if the potential is of short range of nearest neighbor type. The spectrum obtained by Hofstadter shows characteristic structures, and has a deep connection with the structure of the fractional quantum Hall effect. \ (i) Half-filled case, $\nu=1/2$. At half-filling $\nu=1/2$, the system has a half flux $\Phi=\Phi_0/2$. The system, then, is described equivalently with the two-component Dirac field by combining the field at even sites with that at odd sites. In the gauge, $A_x=0,\ A_y=Bx$, the mean field Hamiltonian reads, \begin{eqnarray} &H_{\rm M.F.}= \sum\Psi^\dagger(X') \left( \begin{array}{cc} a_{ee}(X'-X) & a_{eo}(X'-X)\\ a_{oe}(X'-X) & a_{oo}(X'-X) \end{array}\right)\Psi(X), \nonumber\\ &\Psi(X)= \left(\begin{array}{c} c(2m,n)\\ c(2m+1,n) \end{array}\right), \end{eqnarray} \begin{eqnarray} &a_{ee}(m'-m,n'-n)=\langle c^\dagger(2m',n')c(2m,n)\rangle v(2m'-2m,n'-n),\nonumber\\ &a_{oo}(m'-m,n'-n)=\langle c^\dagger(2m+1',n')c(2m+1,n)\rangle v(2m'-2m,n'-n),\nonumber\\ &a_{eo}(m'-m,n'-n)=\langle c^\dagger(2m',n')c(2m+1,n)\rangle v(2m'-2m-1,n'-n),\nonumber\\ &a_{oe}(m'-m,n'-n)=\langle c^\dagger(2m'+1,n')c(2m,n)\rangle v(2m'-2m+1,n'-n),\nonumber\\ &a_{eo}=a_{oe},\\ &a_{ee}+a_{oo}=0.\nonumber \end{eqnarray} We obtain the self-consistent solutions numerically. Fig.2 shows the spectrum. As is expected, spectrum has two minima and two zeros corresponding to doubling. We have the momentum space expression of the mean field Hamiltonian, \begin{eqnarray} &H_{\rm M.F.}&=\sum c^\dagger_\xi({\bf p})F_{\xi\eta}({\bf p})c_\eta ({\bf p})\\ &&=\sum\epsilon_\alpha({\bf p})c^\dagger_\xi({\bf p})U^\dagger _{\xi\xi'}({\bf p})U_{\xi'\eta}({\bf p})c_\eta({\bf p}),\nonumber\\ &&\epsilon_\pm({\bf p})=\pm\sqrt{a_{ee}^2({\bf p})+a_{eo}^2({\bf p})}, \nonumber\\ &&U_{\xi\eta}({\bf p})= \left(\begin{array}{cc} a_{eo}/N_+ & (\epsilon_+ -a_{ee})/N_+\\ a_{eo}/N_- & (\epsilon_- -a_{ee})/N_- \end{array}\right),\nonumber\\ &&N^2_{\pm}=2(a^2_{ee}+a^2_{eo}-a_{ee}\epsilon_{\pm}), \nonumber \end{eqnarray} $$ a_{ee}(p_x,p_y+\pi/a)=-a_{ee}(p_x,p_y),\ a_{ee}(p_y,p_x)=a_{eo}(p_x,p_y).\nonumber $$ The matrix $F_{\xi\eta}({\bf p})$ is approximated well with a nearest neighbor form, \begin{equation} F({\bf p})=2K\left(\begin{array}{cc} \cos p_y&\cos p_x\\ \cos p_x&-\cos p_y \end{array}\right), \end{equation} $$ K=0.107{e^2\over\kappa l_B},\ l_B=\sqrt{\hbar\over eB}. $$ Around minima, the energy eigenvalue of Eq.(3.11) is approximated as, \begin{equation} E(p)=E_0+{({\bf p}-{\bf p}_0)^2\over 2m^*},\ {\bf p}_0=(0,0),(0,\pi/a), \end{equation} and they are approximated around zeros as, \begin{eqnarray} &E(p)=\gamma{\hbox{$\alpha$\kern-.7em\hbox{$\alpha$}}}\cdot{\bf p},\\ &\alpha_x=\left(\begin{array}{cc} 0&1\\ 1&0 \end{array}\right),\ \alpha_y=\left(\begin{array}{cc} 1&0\\ 0&-1 \end{array}\right), \nonumber \end{eqnarray} in $2\times2$ expression. The $m^*$ in Eq.(3.13) is the effective mass and $\gamma$ is the effective velocity. They are computed numerically. Its value is, \begin{eqnarray} &m^*=0.225\sqrt{B\over B_0}m_e,\ B_0=20{\rm Tesla},\\ &\gamma=0.914{e^2\over\kappa}.\nonumber \end{eqnarray} The effective mass is proportional to the square root of $B$ in our method. $\nu=1/2$ mean field Hamiltonian is invariant under a kind of Parity, $P$, and anti-commutes with a chiral transformation, $\alpha_5$, which are defined by, \begin{eqnarray} P:\ &\Psi(p_x,p_y)&\rightarrow \alpha_x\Psi(p_x,p_y+\pi/a),\\ \alpha_5:\ &\Psi(p_x,p_y)&\rightarrow\alpha_5\Psi(p_x,p_y),\nonumber\\ &&\alpha_5=\alpha_x\alpha_y.\nonumber \end{eqnarray} If the parity is not broken spontaneously, there is a degeneracy due to parity doublet. Doubling of the states\cite{r15} appears also at $\nu\neq1/2$ and plays important role, when we discussed the states away from $\nu=1/2$ in the next part. When an additional vector potential with the same gauge, $A_x=0,\ A_y=Bx$, is added, the Hamiltonian satisfies the properties under the above transformations and the doubling due to parity doublet also appears. Thus, the factor 2 is necessary in Eq.(3.3) and leads the principal series at $\nu=p/(2p\pm1)$ to have maximum energy gap. \ (ii) $\nu={p\over 2p\pm1}$ If the filling factor, $\nu$, is slightly away from 1/2, total system can be regarded as a system with a small magnetic field $(\nu-1/2)B_0$. A band structure may be slightly modified. It is worthwhile to start from the band of $\nu=1/2$ as a first approximation and to make iteration in order to obtain self-consistent solutions at arbitrary $\nu=p/(2p\pm1)$. We solve the following mean field Hamiltonian under the self-consistency condition at $\nu=1/2+\delta$, \begin{eqnarray} &H_{\rm M}=\sum U^{({1\over 2}+\delta)}_0({\bf R}_1-{\bf R}_2) e^{i\int({{\bf A}}^{({1\over 2})}+ \delta{\bf A})d{\bf x}}v({\bf R}_1-{\bf R}_2) c^\dagger({\bf R}_1)c({\bf R}_2),\\ &\langle c^\dagger({\bf R}_1)c({\bf R}_2)\rangle_{1/2+\delta}= U^{({1\over 2}+\delta)}_0 e^{i\int({{\bf A}}^{({1\over 2})} +\delta{\bf A})}.\nonumber \end{eqnarray} Here we solve, instead, an Hamiltonian which has the phase of Eq.(3.17) but has the magnitude of the $\nu=1/2$ state. Namely we study a series of states defined at $\nu=1/2$, \begin{eqnarray} H_{\rm M}^{(2)}&=\sum U^{({1\over2})}_0({\bf R}_1-{\bf R}_2) e^{i\int{\bf A}^{({1\over2})}d{\bf x}}v({\bf R}_1-{\bf R}_2) e^{i\int\delta{\bf A}d{\bf x}}c^\dagger({\bf R}_1)c({\bf R}_2) \nonumber\\ &=\sum F^{({1\over2})}({\bf R}_1,{\bf R}_2;\delta{\bf A}) c^\dagger({\bf R}_1)c({\bf R}_2). \end{eqnarray} Integer quantum Hall state has an energy gap of the Landau levels due to $\delta{\bf A}$. This occurs when the integer number of Landau levels are filled completely. Landau level structure is determined by the phase factor and Eq.(3.17) and Eq.(3.18) have the same Landau level structure. Magnitudes of the physical quantities of Eq.(3.17) may be modified, nevertheless. $F^{({1\over2})}({\bf R}_1,{\bf R}_2;0)$ was obtained in the previous part, and is approximated with either the effective mass formula (3.13) or with the nearest neighbor form (3.12). $ $From Eqs.(3.13) and (3.18), the Hamiltonian $H^{(2)}_{\rm M}$ is written in the former method as \begin{eqnarray} H_{\rm M}^{(2)}&=\sum c^\dagger({\bf R}_1)U^\dagger({\bf R}_1, {\bf R}'_1,\delta{\bf A}) \{E_0+{({\bf p}+e\delta{\bf A})^2\over2m^*}\} U({\bf R}'_2,{\bf R}_2,\delta{\bf A})c({\bf R}_2)\nonumber\\ &=\sum\tilde c^\dagger(R_1) \{E_0+{({\bf p}+e\delta{\bf A})^2\over2m^*}\}\tilde c({\bf R}_1),\\ &\tilde c({\bf R}_1)=\sum U({\bf R}_1,{\bf R}'_1)c({\bf R}'_1),\\ &U_{\xi\eta}({\bf R}_1,{\bf R}'_1)=\int d{\bf p} e^{i{\bf p}\cdot({\bf R}_1- {\bf R}'_1)}U_{\xi\eta}({\bf p}). \end{eqnarray} Integer Hall states of Eq.(3.19) satisfy, \begin{equation} \langle{\tilde c}^\dagger_\xi({\bf R}_1) {\tilde c}_\eta({\bf R}_2)\rangle= \tilde u_{\xi\eta}({\bf R}_1-{\bf R}_2) e^{i\int^{{\bf R}_1}_{{\bf R}_2}\delta{\bf A}d{\bf x}}, \end{equation} and leads the expectation value, \begin{eqnarray} \langle c^\dagger({\bf R}_1)c({\bf R}_2)\rangle&=&U^\dagger( {\bf R}_1,{\bf R}'_1) \langle{\tilde c}^\dagger({\bf R}'_1){\tilde c}({\bf R}'_2)\rangle U({\bf R}'_2,{\bf R}_2)\\ &=&U^\dagger({\bf R}_1,{\bf R}'_1)\tilde u_{\xi\eta}({\bf R}'_1-{\bf R}'_2) e^{i\int^{{\bf R}'_1}_{{\bf R}'_2}\delta{\bf A}d{\bf x}} U({\bf R}'_2,{\bf R}_2).\nonumber \end{eqnarray} Here, we solve the equations obtained from Eq.(3.19) with continuum approximation, first. We have, \begin{eqnarray} &[E_0+{({\bf p}-{\bf p}_0^{(i)}+\delta{\bf A}_{\rm ind})^2\over 2m^*}] u^{(i)}_p=E_p u_p^{(i)},\\ &i=1,2,\nonumber\\ &E_p=E_0+{e\delta B_{\rm eff}\over 2m^*}(2p+1)\\ &\delta B_{\rm eff}=\vert \nu-{1\over2}\vert B_0.\nonumber \end{eqnarray} $ $From Eq.(3.25), $p$ Landau levels are completely filled and the integer quantum Hall effect occurs at $\nu=p/(2p\pm1)$. The energy gap is given by Landau level spacing, \begin{equation} \Delta E_{\rm gap}={e\delta B_{\rm eff}\over m^*}={eB_0\over m^*} \vert\nu-{1\over2}\vert. \end{equation} Hamiltonian with the nearest neighbor form is a $2\times2$ matrix of the tight-binding type. Based on this Hamiltonian, we solve Landau level equation of the tight-binding form with the vector potential $\delta{\bf A}$, \begin{eqnarray} &K\{c_1({\bf n}+{\bf\Delta}_x)e^{i\int_{\bf n}^{{\bf n}+{\bf\Delta}_x} \delta{\bf A}\cdot d{\bf x}}+ c_2({\bf n}+{\bf \Delta}_y)e^{i\int_{\bf n}^{{\bf n}+{\bf\Delta}_y} \delta{\bf A}\cdot d{\bf x}}+{\rm h.c.}\}=Ec_1({\bf n}),\nonumber\\ &K\{c_1({\bf n}+{\bf\Delta}_x)e^{i\int_{{\bf n}}^{{\bf n}+{\bf\Delta}_x} \delta{\bf A}\cdot d{\bf x}}- c_2({\bf n}+{\bf\Delta}_y)e^{i\int_{\bf n}^{{\bf n}+{\bf\Delta}_y} \delta{\bf A}\cdot d{\bf x}}+{\rm h.c.}\}=Ec_2({\bf n}).\nonumber\\ & \end{eqnarray} The equation are solved numerically and the energy gaps and the widths of excited bands are given in Fig.3. Some bands are narrow and some bands are wide. Near $\nu=1/2$, the effective magnetic field approaches to zero and Landau level wave functions have large spatial extensions. Lattice structure becomes negligible and spectrum shows simple Landau levels of the continuum equation in these regions. Near $\nu=1/3$, lattice structure is not negligible and bands have finite widths. There are non-negligible corrections from those of continuum calculations, Eq.(3.24). Due to energy gap of the integer Hall effect caused by the induced dynamical magnetic field, the states at $\nu=p/(2p\pm1)$ are stable and fluctuations are weak. Invariance under $P$, moreover, ensures these states to have uniform density. In systems with impurities, localized states with isolated discrete energies are generated by impurities and have energies in the gap regions. They contribute to the density but do not contribute to the conductance. If the Fermi energy is in one of these gap regions, the Hall conductance $\sigma_{xy}$ is given by a topological formula, Eq.(2.19), and stays constant, at ${e^2\over h}\cdot{p\over 2p\pm1}$. The fractional Hall effect is realized. At a value of $\nu$ smaller than 1/3, Hofstadter butterfly shows other kind of structures hence suggests other structures of the fractional Hall effect do exist. For instance at $\nu=$1/5, 1/7, 1/9,$\dots$the energy gap is large. In low density, it may be complicated, in fact. There could exist completely different kind of phase, such as Wigner crystal phase\cite{r16}. Competition between two phases may be important. They will not be presented in the present paper but it will be presented in a later work. \ (iii) Fluctuations of FQHE at $\nu={p\over 2p\pm1}$. Ground states have energy gap, hence the fluctuations are small, just as in the integer Hall effect. Density and phase fluctuations are described by the massive Chern-Simons gauge theory\cite{r17}with a mass of order energy gap. \ (iv) Fluctuations at half-filling, $\nu={1\over2}$. Ground state has no energy gap at $\nu=1/2$. Fluctuations are described by the action, \begin{equation} S=\int d^3x\Psi^\dagger(i\hbar{\partial\over\partial t}+a_0)\Psi+ \gamma\Psi^\dagger\hbox{$\alpha$\kern-.7em\hbox{$\alpha$}}\cdot({\bf p}+{\bf a})\Psi \end{equation} The fermion field integration leads severe infra-red divergence. If the energy dispersion is changed to \begin{equation} E_0=\tilde\gamma\vert p\vert^{1+\delta}, \end{equation} by interaction, physics at $\nu=1/2$ is completely different from that of the mean field. We will not discuss physical properties of $\nu=1/2$ state in this paper. \section{Comparison with experiments}\setcounter{equation}{0} In the previous section we presented our mean field theory based on flux condensation, where lattice structure generated by the external magnetic field and condensed flux due to interaction are important ingredient. Consequently, our mean field Hamiltonian becomes very similar to that of Hofstadter which is known to show actually large energy gap zone along $\Phi=\nu\Phi_0$ line. The line $\Phi=\nu\Phi_0$ is special in Hofstadter problem and hence in our mean field Hamiltonian, too. This explains why the experiments of the fractional quantum Hall effects shows characteristic behavior at $\nu=p/(2p\pm1)$. The ground states at $\nu=p/(2p\pm1)$ have lowest energy and largest energy gap, hence these states are stable. In this section we compare the energy gaps of the principal series with the experiments in the lowest order. and the ground state energy of $\nu=1/3$ state with the Laughlin variatioal wave function\cite{r18}. The effective mass $m^*$ of Eq.(3.14) was obtained from the curvature of the energy dispersion and should show a characteristic mass scale of the fractional Hall effect. Eq.(3.24) gives Landau level energy in the lowest approximation and the gap energy is given in Eq.(3.25). The gap energy from the nearest neighbor approximation, Eq.(3.12) is given by solving Eq.(3.27). They are compared with the experimental values\cite{r19} in Fig.4. The agreement is not perfect but should be regarded good as the lowest mean field approximation. Near $\nu=1/3$, the bands have finite widths and near $\nu=1/2$, the widths are infinitesimal. The dependence of the width upon the filling factor, $\nu$, and the whole structure of the bands expressed in Fig.1 are characteristic features of the present mean field and should be tested experimentally. Finally we compute the ground state energy of the $\nu=1/2$ state and the $\nu=1/3$ state. Using the fact that our mean field Hamiltonian of $\nu=1/3$ is very close to the short-range tight-binding Hamiltonian, we compute the ground state energy of $\nu=1/3$ with the tight-binding model wave function. The wave function is obtained numerically and is substituted to the total energy per particle, Eq.(3.7), as \begin{eqnarray} E_{1/3}&=&{3\over N}\langle\Psi\vert H\vert\Psi\rangle \\ &=&-{3\over2}\sum_{\bf X}v({\bf X})\vert U_0({\bf X})\vert^2 .\nonumber \end{eqnarray} where $N$ is the number of sites. The result is, \begin{equation} E_{1/3}=-0.340{e^2\over\kappa l_B}. \end{equation} where $l_B=\sqrt{\hbar/eB}$. This value should be compared with that of Laughlin wave function, \begin{equation} E_{1/3}=-0.416{e^2\over\kappa l_B}. \end{equation} The value of Eq.(4.2) is higher than Eq.(4.3), but the difference is not large. This may suggest that mean field flux state is infact close to Laughlin wave function and to the exact solution. For the state at $\nu=1/2$, we use our self-consistent solution for computing the ground state energy per particle, \begin{equation} E_{1/2}=-0.347{e^2\over\kappa l_B}. \end{equation} As is mentioned in Section.3, the fluctuations are extremely large and the mean field value get large corrections from the higher order effects. So the results about the $\nu=1/2$ state should not be taken seriously. \section{Summary}\setcounter{equation}{0} We formulated the quantum Hall effects, integer Hall effect and fractional Hall effect with von Neumann lattice representation of two-dimensional electrons in a strong magnetic field. von Neumann lattice is a subset of coherent state. The overlapp of the states is expressed with theta function. They give a systematic way of expressing quantum Hall dynamics. Topological invariant expression of the Hall conductance was obtained in which compactness of the momentum space is ensured by the lattice of the coordinate space. Because the lattice has an origin in the external magnetic field, topological character of the Hall conductance is ensured by the external magnetic field. The conductance is quantized exactly as $(e^2/h)\cdot N$ at the quantum Hall regime. The integer $N$ increases with chemical potential. A new mean field theory of the fractional Hall effect that has dynamical flux condensation is studied. If the filling $\nu$ is less than one, many particle states of Landau levels have no energy gap unless interaction switchs on. In the tight-binding model, the situation is very different. The spectrum that was found by Hofstadter first is changed drastically when the flux is changed and it has a large energy gap in some regions. In our mean field theory, lattice structure is introduced from von Neumann lattice and flux is introduced dynamically. The mean field Hamiltonian becomes a kind of tight-binding model and rich structure of the tight-binding model is seen, in fact, as characteristic features of the fractional Hall effect. Our mean field flux states have liquid property of uniform density with energy gap. They are defined as special integer quantum Hall states in the lowest Landau level space hence the band structures shown in Figs.2 and 3 are very different from those of normal integer Hall states. We gave a dynamical reason why the principal series at $\nu=p/(2p\pm1)$ are observed dominantly. These states satisfy the self-consistency condition of having the lowest energy and the largest energy gap. The physical quantities of our mean field theory are close to the experimental values in the lowest order at $\nu=p/(2p\pm1)$. At the exact half-filling $\nu=1/2$, fluctuations are very large and corrections from mean field value may be large, too. \section*{Acknowledgements} We are indebted to Professors P. Wiegmann and H. Suzuki for their useful comments in the early stage of the present work. One of the authors(K.I.) thanks Professors A. Luther and H. Nielsen for useful discussions. The present work was partially supported by the special Grant-in-Aid for Promotion of Education and Science in Hokkaido University Provided by the Ministry of Education, Science, Sports, and Culture, a Grant-in-Aid for Scientific Research(07640522), and Grant-in-Aid for International Scientific Research(Joint Research 07044048) the Ministry of Education, Science, Sports, and Culture, Japan. \newpage \renewcommand{\theequation}{A.\arabic{equation}} \noindent {\Large \bf APPENDIX A} \setcounter{equation}{0} \bigskip In the present representation, a short-range impurity term can be expressed as, \begin{eqnarray} &H_{\rm impurity}=a^4\int^{\pi/a}_{-\pi/a}{d^2p_1\over (2\pi)^2} {d^2p_2\over (2\pi)^2}b^\dagger(p_1)\sum_{{\bf N}}({2\pi\over a})^2 e^{ia\int^{p_2^N}_{p_1}{\bf A}\cdot d{\bf p}}\tilde V(p_1-p_2^N)e^{ -{a^2(p_1-p_2^N)^2\over8\pi}}b(p_2^N)\nonumber\\ &=a^2\int^{\pi/a}_{-\pi/a}{d^2p_1\over (2\pi)^2}b^\dagger(p_1)\int^\infty _{-\infty}d^2 k\tilde V(-k)e^{-{a^2k^2\over8\pi}}e^{iak^i D_i}b(p_1), \nonumber\\ &D_i={1\over i}{\partial\over a\partial p_i}+A_i,\ {\bf A}=({ap_y\over2\pi},0), \\ &{\bf p}_2^N={\bf p}_2+2\pi{\bf N},\ V(x)=\int^\infty_{-\infty}d^2k\tilde V(k)e^{i{\bf k}\cdot{\bf x}}. \nonumber \end{eqnarray} With the creation and annihilation operator defined by, \begin{eqnarray} &A=D_x-iD_y,\\ &A^\dagger=D_x+iD_y,\nonumber\\ &[A,A^\dagger]={1\over\pi},\nonumber \end{eqnarray} the above Hamiltonian can be written as, \begin{equation} H_{\rm impurity}=a^2\int^{\pi/a}_{-\pi/a}{d^2p\over (2\pi)^2} b^\dagger(p)\int^\infty_{-\infty}d^2 k\tilde V(-k)e^{{a\over2}(ik_x-k_y)A} e^{{a\over2}(ik_x+k_y)A^\dagger}b(p). \end{equation} We represent $H_{\rm impurity}$ in Landau level representation of the momentum space defined by, \begin{eqnarray} b(p)&=&\sum_l b_l\tilde\psi_l(p),\ \vert l\rangle=\pi^{l\over2}{1\over \sqrt{l!}} (A^\dagger)^l\vert 0\rangle,\\ A\vert 0\rangle&=&0,\nonumber\\ \tilde\psi_l(p)&=&\langle p\vert l\rangle =C\sum_N e^{-i\pi N}e^{ik_N a p_x}H_l({a(p_y+{2\pi\over a} k_N)\over\sqrt{\pi}})e^{-{a^2\over4\pi}(p_y+{2\pi\over a} k_N)^2}, \nonumber \end{eqnarray} where $k_N=N+{1\over2}$ and $C$ is the normalization constant. The normalized lowest Landau level wave function is, \begin{equation} \tilde\Psi_0=2^{1/4}e^{-{a^2 p_y^2\over4\pi}} \Theta_1({a(p_x+ip_y)\over2\pi},i). \end{equation} $H_{\rm impurity}$ is reduced to, \begin{eqnarray} &H_{\rm impurity}=\sum_{l_1,l_2}b^\dagger_{l_1}\tilde V_{l_1,l_2} b_{l_2},\\ &\tilde V_{l+n,l}=4\pi\sqrt{l!\over (l+n)!}\int d^2 k \tilde V(-2\sqrt{\pi} k)a^n (ik_x+k_y)^n L_l^{(n)} (k^2 a^2)e^{-a^2 k^2},\nonumber\\ &\tilde V_{l,l+n}=4\pi\sqrt{l!\over (l+n)!}\int d^2 k \tilde V(-2\sqrt{\pi} k)a^n (ik_x-k_y)^n L_l^{(n)} (k^2 a^2)e^{-a^2 k^2}.\nonumber \end{eqnarray} For a short range potential, we have \begin{eqnarray} &V(x)=g\delta(x),\ \tilde V(-k)={g\over (2\pi)^2},\nonumber\\ &\tilde V_{l+n,l}=\tilde V_{l,l+n}=0,\ {\rm for}\ n\neq0,\\ &\tilde V_{l,l}=0,\ {\rm for}\ l\neq0,\nonumber\\ &\tilde V_{0,0}={g\over (2\pi)^2}4\pi\int d^2 k L_0^{(0)}(a^2k^2)e^{-a^2k^2}={g\over a^2},\nonumber\\ &H_{\rm impurity}=\tilde V_{0,0}b^\dagger_0 b_0. \end{eqnarray} Hence the one Landau level has the energy shift $\tilde V_{00}$, and all the other Landau levels have no effect from the impurity. The state $\vert l=0\rangle$ has an isolated energy and its wave function is square-integrable. This state corresponds to localized state. \newpage \renewcommand{\theequation}{B.\arabic{equation}} \noindent {\Large \bf APPENDIX B} \setcounter{equation}{0} \bigskip In Chern-Simons gauge theory in momentum space, the action is \begin{equation} \int dtd{\bf p} {1\over2}\epsilon_{\mu\nu\rho}a^\mu{\partial\over\partial p_\nu} a^\rho+a_0 j^0({\bf p})+ F(c^\dagger(p_1)e^{i\tilde e\int^{p_1}_{p_2}a_i dp_i}c (p_2)), \end{equation} where in the last term operators $b^\dagger(p_1)b(p_2)$ are replaced with $c^\dagger(p_1)e^{i\tilde e\int^{p_1}_{p_2}a_i dp_i}c (p_2)$, and the theory is defined on the torus\cite{r20}. The vector potential $a_i$ satisfy, \begin{equation} \epsilon_{0ij}{\partial\over\partial p_i} a^j+j^0({\bf p})=0. \end{equation} Its solution is substituted into the first term of (B.1). By choosing the coupling strength $\tilde e$ from the condition, \begin{eqnarray} \tilde e j^0({\bf p})+{a^2\over 2\pi}=0,\\ j^0({\bf p})=({a\over 2\pi})^2 N_{\rm total}. \end{eqnarray} the phase factor which expresses the magnetic field in the momentum space in Eqs.(2.11), (2.12), (2.13), and (2.14) are cancelled. In thermodynamic limit, $j^0({\bf p})$ diverges, hence $\tilde e$ in infinitesimally small. The fluctuation of Chern-Simons gauge field can be ignored. The first two terms in Eq.(B.1) cancell if Eq.(B.2) is satisfied and do not contribute to the energy of the system. \newpage
proofpile-arXiv_065-435
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} During its 17 plus years of operation, the {\it International Ultraviolet Explorer\/} ({\it IUE\/}) satellite (\cite{bog78}) acquired a large number ($> 10^5$) of UV spectra from all classes of astronomical objects. Cataclysmic variables have been studied intensively with {\it IUE\/}, and these spectra have revealed a wide variety of phenomena which were previously unknown or poorly studied. These include mass loss via high-velocity winds, the heating and cooling of the white dwarf, the spectrum of the accretion disk, and the response of the disk to dwarf nova outbursts. As a result, we now have a much more complete understanding of the nature of CVs: the geometry, velocity law, mass-loss rate, and ionization structure of the wind; the disk instability mechanism and the heating and cooling waves which drive the disk to and from outburst; the relative energetics of the accretion disk and boundary layer; subclasses; and even evolutionary scenarios. These advances are due in part to the fact that CVs emit a large fraction of their luminosity in the UV. With few exceptions, the strategy which has prevailed to date in the study of CVs with {\it IUE\/} and other instruments is to make an observation or set of observations of a single object and to analyze and interpret those data separately from the data of other similar objects. Only when this step is complete are the data compared with those from other similar objects. An alternative strategy is to examine the data from all the objects of a given class simultaneously. This is practical only in situations where a large body of data exists, of fairly uniform quality, and when the various members of a given class form a relatively homogeneous group. A few such studies of CVs using various manifestations of the {\it IUE\/} archive have appeared in the literature. Verbunt (1987) studied the absolute continuum spectral flux distributions of disk-fed CVs. la Dous (1991) studied the relative continuum spectral flux distributions of all classes of CVs, but concentrated on disk-fed systems. For the discrete spectral features, only the inclination dependence of the equivalent widths was discussed. Deng et~al.\ (1994) studied the dependence on the orbital period and inclination of the relative continuum spectral flux distributions and the equivalent widths of the lines of dwarf novae in quiescence. It is unfortunate that so little quantitative work has been performed to date on the emission lines of CVs. Emission lines are powerful diagnostics of the physical conditions in all types of objects including these interacting binaries. The emission line fluxes or relative fluxes can constrain the temperature, ionization state, density, elemental abundance, and geometrical distribution of the gas. However, the dependence of measurable quantities on these parameters is often complicated, necessitating numerical model fitting in many situations. The range of possible assumptions regarding the physical mechanisms of heating and ionization, together with the technical challenges of constructing models, add to the difficulty of performing such studies. As a result, emission lines have received less attention as diagnostics of CVs than have continuum properties. In the present paper, we begin to remedy this situation with an analysis of the flux ratios of the emission lines of all classes of CVs using spectra extracted from the {\it IUE\/} Uniform Low Dispersion Archive (ULDA). We begin by describing the data and our analysis procedures (\S 2). We then present the results of the data analysis (\S 3), followed by a discussion of models for the line emission (\S 4). In \S 5 we compare data and models and in \S 6 close with a summary of our conclusions. \section{Data Analysis} The {\it IUE\/} ULDA (\cite{wam89}) provides a unique source of archival data which satisfies the requirements for a comprehensive study of a large sample of objects owing to its uniform quality and the large number of spectra accumulated over the long life of {\it IUE\/}. Although the {\it IUE\/} final archive (\cite{nic94}) has since superseded the ULDA, the reprocessing effort of the final archive had not begun when this study was begun in 1992. To build the data sample, the {\it IUE\/} Merged Log was searched using software at the {\it IUE\/} Regional Analysis Facility, then at the University of Colorado. The search criteria were the camera (SWP), dispersion (low), and {\it IUE\/} object class (dwarf novae, classic novae, irregular variables, and nova-like variables). All spectra meeting these criteria were extracted from the ULDA resident at Goddard Space Flight Center, which at that time contained (nearly) all low-dispersion {\it IUE\/} spectra through 1988 (SWP sequence numbers through $\approx 35190$). The spectra were then written to disk with header information extracted from the {\it IUE\/} Merged Log. Excluded from further consideration were 13 very early (prior to 1978 July) spectra of AM~Her and SS~Cyg lacking data quality flags. The final dataset consisted of $\approx 1300$ spectra of $\approx 100$ CVs. From this dataset, we selected all spectra exhibiting pure emission lines. Sources which met this criteria include AM~Her stars, DQ~Her stars, dwarf novae in quiescence, and eclipsing nova-like variables and dwarf novae in outburst. To this dataset we added the {\it IUE\/} spectra of the eclipsing nova-like variable V374~Pup listed in Table~1 of Mauche et~al.\ (1994), as we were working on that source at the time. Excluded from analysis were non-eclipsing nova-like variables and dwarf novae in outburst, since the UV resonance lines of such systems are either in absorption or are P~Cygni profiles. For these spectra, we measured the integrated fluxes above the continuum of the strongest emission lines in the SWP bandpass: N~V $\lambda 1240$, Si~IV $\lambda 1400$, C~IV $\lambda 1550$, and He~II $\lambda 1640$. To determine the flux of these lines, the continuum to the left and right of each line (e.g., from 1500--1530 \AA \ and from 1570--1600 \AA \ for C~IV) was fitted with a linear function by the method of least squares. This method works well with all of the emission lines except N~V, because the left side of this line falls on the wing of the geocoronal Lyman $\alpha $ emission line. To overcome this problem, a continuum point to the left of the N~V line was specified interactively, and the fit included the region to the right of the line (1255--1270 \AA ) plus this continuum point to the left. Care was taken to insure that the continuum over this small wavelength region was a reasonable extrapolation both in normalization and slope to the continuum at longer wavelengths. In all cases the least-squares fit was made under the assumption that the errors $\sigma _i$ in the flux density $f_i$ at each wavelength point $\lambda _i$ were the same. After the normalization $a$ and slope $b$ of the continuum was determined, the size of the errors $\sigma $ were estimated by assuming that the reduced $\chi ^2$ of the fit was equal to 1: $\sigma ^2 = \sum _{i=1}^N (f_i-a-b\lambda _i)^2/(N-2)$. This procedure is necessary because the ULDA (unlike the {\it IUE\/} final archive) does not include the error associated with each flux point. The integrated fluxes of the emission lines were then determined by summing the flux minus the fitted continuum over the region of the line (e.g., 1530--1570 \AA \ for C~IV). The error on the integrated flux was determined by propagating the error $\sigma $ through the calculation. Finally, careful track was kept of the number of pixels labeled by the data quality flags as either saturated or nearly saturated (extrapolated ITF). Such pixels compromise the quality of the flux density measurements and consequently the integrated line fluxes; saturated pixels are particularly bad in this regard, because they systematically produce only lower limits to the flux density. In the final cut to the data, flux measurements for a given line were retained only if the total number of saturated and extrapolated pixels included in the summation was less than or equal to some number $n$. While $n=1$ would have been the ideal choice, such a stringent selection criterion would cull too many measurements. As a reasonable compromise, we settled on $n=3$ for this selection criterion, allowing a few mild sinners to pass with the saints. The roster at this stage of the analysis consisted of $\approx 700$ spectra of $\approx 60$ CVs. Of this number, we present results based on $\approx 430$ spectra of the 20 systems with the largest number ($> 10$) of flux measurements. \section{Observational Results} Just as color-color diagrams are a useful way to discuss the continuum spectral flux distributions of CVs (e.g., \cite{wad88}; \cite{den94}), flux ratio diagrams are a useful way to display and discuss the fluxes of their emission lines. Flux ratios remove the variables of luminosity and distance from the comparison of different sources; flux ratios are unaffected by aperture losses, which occur for {\it IUE\/} spectra obtained through the small aperture; and flux ratios are much less sensitive to reddening than the fluxes themselves. Due to the increase of the extinction efficiency at short wavelengths, reddening has the largest differential effect on the fluxes of the He~II and N~V lines. Luckily, for a typical extinction of $E_{B-V}\leq 0.1$ (\cite{ver87}), the ratio of the measured fluxes of these lines is only $\leq 18$\% higher than intrinsic ratio. As we shall see, this correction, while systematic, is much smaller than either the typical measurement errors, or the typical dispersion in the flux ratios of any given source. Of the many possibilities, we choose to plot the ratios N~V $\lambda 1240$/C~IV $\lambda 1550$ and He~II $\lambda 1640$/C~IV $\lambda 1550$ vs.\ Si~IV $\lambda 1400$/C~IV $\lambda 1550$ (Figs.~1--8). In what follows, we often refer to these ratios simply as N~V/C~IV and He~II/C~IV vs.\ Si~IV/C~IV. It should be emphasized that the identifications of these lines are not unique; at $\Delta\lambda\approx 5$~\AA , the {\it IUE\/} resolution is not sufficient to exclude several other possible lines in the vicinity of these wavelengths. We defer a discussion of such potential confusion, and of the physical motivation for our choice of lines until the following section, and begin by presenting the phenomenological behavior of these ratios as derived from the data. We begin by discussing the various members of the various CV subclasses and then discuss the sources which do not appear to fit the pattern established by the other members of their subclass (sources we enjoy referring to as ``weird''). \subsection{Results Arranged by Subclass} \subsubsection{AM~Her stars: AM~Her, V834~Cen, and QQ~Vul} Figure~1 shows that AM~Her, V834~Cen, and QQ~Vul form a sequence of increasing N~V/C~IV with increasing Si~IV/C~IV. He~II/C~IV is reasonable constant at $\approx 0.2\pm 0.1$. \subsubsection{DQ~Her stars: EX~Hya, TV~Col, FO~Aqr, and DQ~Her} Among these DQ~Her stars, TV~Col displays the greatest amount of variability in its continuum and line fluxes, and, as is evident in Figure~2, in its UV line ratios. The upper right portion of the observed range of line ratios is populated by the observations obtained during the optical and UV flare observed by Szkody \& Mateo (1984); during the flare, both N~V and He~II increased relative to C~IV. The relatively tight phase space occupied by the line ratios of FO~Aqr is approximately the same as that of TV~Col in outburst. The ionization state of the line-emitting gas in EX~Hya and DQ~Her is lower than the two other DQ~Hers, as evidenced by their higher Si~IV/C~IV line ratios and the lower He~II/C~IV line ratios. \subsubsection{Dwarf Novae in Quiescence: SS~Cyg, SU~UMa, RU~Peg, and WX~Hya} As shown in Figure~3, WX~Hyi has the largest ratios of N~V/C~IV and Si~IV/C~IV among all the AM~Her stars, DQ~Her stars, and other dwarf novae in quiescence. SU~UMa and RU~Peg have N~V/C~IV ratios of $\approx 0.3$, but the value for SS~Cyg is $\approx 30\%$ lower at $\approx 0.2$. \subsubsection{Eclipsing Dwarf Novae: Z~Cha and OY~Car} As shown in Figure~4, relative to all other ``normal'' CVs, Z~Cha and OY~Car both have strong N~V relative to C~IV. Almost all of the spectra of Z~Cha were obtained during the peak and decline of the superoutburst of 1987 April (\cite{har2a}) and of a normal outburst of 1988 January (\cite{har2b}). The evolution of the spectrum during these observations is manifest most clearly in the Si~IV/C~IV ratio, which was highest during the decline of the superoutburst ($V\approx 12.3$--12.7) and lowest during the decline of the normal outburst ($V\approx 13.3$--14.0). During these intervals, both the N~V/C~IV and He~II/C~IV ratios remained roughly constant, though there is some indication that the He~II/C~IV ratio increased as the Si~IV/C~IV ratio decreased. The spectra of OY~Car were obtained during the decline of the superoutburst of 1985 May (\cite{nay88}) and produce flux ratios which roughly equal those of Z~Cha in superoutburst. \subsubsection{Eclipsing Nova-like Variables: UX~UMa, V347~Pup, and RW~Tri} As shown in Figure~5, the eclipsing nova-like variables UX~UMa, V347~Pup, and RW~Tri have moderately strong N~V relative to C~IV, with N~V/C~IV ratios intermediate between those of dwarf novae in quiescence and eclipsing dwarf novae. The phase space occupied by V347~Pup is amazingly tight, and is disturbed only by eclipse effects: the single discrepant Si~IV/C~IV ratio was obtained from the eclipse spectrum of this source, and is due to that fact that the C~IV emission line is eclipsed less than the other emission lines. The cause of the large spread in line ratios of UX~UMa appears to due to a different mechanism: there are epochs when Si~IV is weak relative to C~IV, and other epochs when it is reasonably strong. These epochs do not seem to correlate with the continuum flux. The He~II/Si~IV ratio of UX~UMa is typically less than that of V347~Pup. \subsubsection{Intercomparison of Various Subclasses} Figure~6 combines the line ratios of the objects from the previous figures grouped together into magnetic (upper panels) and non-magnetic (lower panels) systems. It is apparent from this figure and the previous discussion that the line ratios of the various subclasses are rather homogeneous, with a dispersion of $\sim 1$ decade. This dispersion to be due to almost equal contributions from the dispersion of values from the various observations of individual objects ($\sim 0.5$ decade) and from the dispersion from one object to another ($\sim 0.5$ decade). There are clear and significant differences between these two subclasses: (i) He~II/C~IV is $\sim 0.25$ decades larger in magnetic systems; (ii) Si~IV/C~IV is slightly larger in the non-magnetic systems, but the difference is small compared to the dispersion; (iii) The upper limit on the N~V/C~IV distribution is greater in the non-magnetic systems, while the lower limit of the distribution is similar in the two cases. It is also instructive to consider the differences within these two subclasses. The greater range of the N~V/C~IV distribution is due to the eclipsing dwarf novae Z~Cha and OY~Car (squares and circles, respectively, in the lower panels of Fig.~6); otherwise the N~V/C~IV distributions of magnetic and non-magnetic systems are similar. Among the magnetic systems, TV Col (triangles in the upper panels of Fig.~6) has significantly lower Si~IV/C~IV than any other object; the Si~IV/C~IV distribution would be much tighter if this object were excluded from the sample. \subsection{Systems Showing Anomalous Behavior: ``Weird'' CVs} \subsubsection{GK Per} GK~Per is unusual among CVs in a number of respects. It has an evolved secondary and hence a very long orbital period and a large accretion disk. Unlike most CVs, the UV continuum of this DQ~Her-type system peaks in the {\it IUE\/} bandpass. The cause of this anomaly is thought to be due to the disruption of the inner disk by the magnetic field of the white dwarf (\cite{bia83}; \cite{wu89}; \cite{mau90}; \cite{kim92}). As shown in Figure~7, GK~Per also distinguishes itself from other CVs in its anomalously large He~II/C~IV ratio. In outburst, both He~II and N~V are stronger than C~IV, whereas the opposite is true in quiescence. While the Si~IV/C~IV ratio spans a rather broad range of $\approx 0.2$--1, this range is present in both outburst and quiescence. \subsubsection{V~Sge} V~Sge is an unusual nova-like variable which occasionally manifests brightenings of as much as 3 mag (\cite{her65}). The model for this system is highly uncertain; both white dwarf and neutron star binary models have been considered (\cite{koc86}; \cite{wil86}). The {\it IUE\/} spectra of V~Sge obtained during 1978 and 1979 are characterized by strong He~II and N~V emission lines relative to C~IV (\cite{koc86}): N~V/C~IV $\approx 2.5$ and He~II/C~IV $\approx 3$. These distinguishing line ratios are suppressed in outburst. In 1985 August, the UV continuum was enhanced by factor of $\approx 2$ relative to the earlier spectra, while the N~V/C~IV ratio fell to $\approx 0.4$ and the He~II/C~IV ratio fell to $\approx 1$. In 1985 April, the UV continuum was enhanced by factor of $\approx 4$, while the N~V/C~IV ratio fell to $\approx 0.15$; measurements of the He~II/C~IV ratio are not possible during this epoch because the He~II emission line was overexposed. This behavior is opposite to that of GK~Per, whose N~V/C~IV and He~II/C~IV ratios increased mildly in outburst. \subsubsection{BY~Cam} BY~Cam (H~0538+608) is one of only three AM~Her stars above the period gap (V1500 Cyg and AM~Her being the other two) and one of only two which rotates asynchronously (V1500 Cyg being the other one). As pointed out by Bonnet-Bidaud \& Mouchet (1987), BY~Cam is unusual in having weak C~IV and strong N~V. Figure~7 shows that all the line ratios are large because C~IV is weak. \subsubsection{AE~Aqr} AE~Aqr exhibits a host of bizarre phenomena including variable flare-like radio and optical emission; rapid, coherent oscillations; and QPOs. Its UV spectrum has been discussed by Jameson et~al.\ (1980) and is distinguished by the near-absence of C~IV. Unfortunately, the flux measurements of the C~IV emission line are very uncertain because this line is so weak and because the measurements are possibly contaminated by the Si~II 3p--4s multiplet at 1526.71, 1533.43~\AA \ (e.g., \cite{kel87}). Nevertheless, it is clear that N~V and Si~IV have similar strengths and are $\approx 10$ times stronger than He~II and C~IV, which have similar strengths. \section{Line Ratio Modeling} We now present some simple models for the systematic behavior we expect from CV line ratios. Although CVs as a class share a number of fundamental properties, they also differ between subclasses in the likely properties of the gas responsible for line emission. For example, in disk-fed systems, such as nova-like variables and dwarf novae, a possible site for line emission is the disk atmosphere. In this case, the atmosphere may be heated either by viscosity (e.g., \cite{sha91}) or by photons from other parts of the disk or from the vicinity of the white dwarf (e.g., \cite{ko96}). A wind is also present in nova-like variables and dwarf novae in outburst (see \cite{mau6b} for a recent review of CV winds). In AM~Her stars, the line emission may come from the stream of material being transferred from the secondary to the white dwarf which is heated either mechanically or by photons from the shock at the white dwarf surface. DQ~Her stars may have emission from either or both of these sites. Beyond these simple considerations, there is considerable uncertainty about the nature of the emission region: the geometrical arrangement of the gas relative to the source of ionizing photons (if present), the density and optical depth of the gas, the elemental abundances of the gas, the spectrum and flux of the ionizing radiation, and the presence and distribution of any other sources of heating. These quantities are not only uncertain, they may differ from one object to another in a subclass. Furthermore, all subclasses contain anomalous members. We therefore consider models for CV emission regions which employ the very simplest assumptions, namely, slab models each with a single ionizing spectrum, gas composition, and density. We then explore plausible choices for photon flux, column density, and mechanical heating rate (if any). The results serve not only as a test of the validity of the models when compared to the observations, but also provide insight into the physics of the line emission which will hopefully remain useful when more detailed models are considered. Further support for this strategy comes from the results of the previous section: the objects of a given subclass show significant clustering in the two line ratio diagrams, and there are clear differences from one subclass to another. Therefore, we can define our modeling goals as follows: to test whether one or more sets of models can reproduce the mean values of the observed line ratios for the various subclasses of objects, and to test whether any models can reproduce the dispersion or systematic variability of the observed line ratios. Given the uncertainties of the models and the observed dispersion in line ratios, we will consider agreement between the models and observations to be adequate if the two agree to within the dispersion of the observed values, i.e., approximately one decade. We will show that although this is clearly a very crude criterion, it turns out to be very constraining for the models. \subsection{Model Ingredients} Although the mechanism of line emission is uncertain, we favor photoionization as the source of ionization, excitation, and heating of the gas. This is partly because the observed continuum spectra, and their extrapolation into the unobservable spectral regions, provide a convenient and plausible energy source. Further support for this idea comes from Jameson et~al.\ (1980) and King et~al.\ (1983) who compared the observed line strengths from AE Aqr and UX UMa to simple predictions of both photoionized and coronal models and found the coronal models inconsistent with the observed lines. Nevertheless, for the sake of completeness, we examine both photoionized and coronal models. The models are calculated using the XSTAR v1.19 photoionization code (\cite{kal82}; \cite{kal93}). The models consist of a spherical shell of gas with a point source of continuum radiation at the center; this may be used to represent a slab in the limit that the shell thickness is small compared with the radius. In what follows we will use the line fluxes emergent from the illuminated face of the slab, which is equivalent to those emitted into the interior of the spherical shell. The input parameters include the source spectrum, the gas composition and density, the initial ionization parameter (determining the initial radius, see below for a definition), and the column density of the shell (determining the outer radius). Construction of a model consists of the simultaneous determination of the state of the gas and the radiation field as a function of distance from the source. The state of the gas at each radius follows from the assumption of a stationary local balance between heating and cooling and between ionization and recombination. When the gas is optically thin, the radiation field at each radius is determined simply by geometrical dilution of the given source spectrum. Then, as shown by Tarter et~al.\ (1969), the state of the gas depends only on the ionization parameter $\xi$, which is proportional to the ratio of the radiation flux to the gas density. We adopt the definition of the ionization parameter used by Tarter et~al.: $\xi=L/(nr^2)$, where $L$ is the ionizing energy luminosity of the central source (between 1 and 1000 Ry), $n$ is the gas density, and $r$ is the distance from the source. This scaling law allows the results of one model calculation to be applied to a wide variety of situations. For a given choice of spectral shape, this parameter is proportional to the various other customary ionization parameter definitions: $U_H=F_H/n$, where $F_H$ is the incident photon number flux above the hydrogen Lyman limit; to $\Gamma= F_\nu(\nu_L)/(2hcn)$, where $F_\nu(\nu_L)$ is incident energy flux at the Lyman limit; and to $\Xi=L/(4\pi R^2 cnkT)$. This simple picture breaks down when the cloud optical depth is non-negligible, since the source spectrum then depends on position, and the escape of cooling radiation in lines and recombination continua depends on the total column density of each ion species and hence on the ionization state of the gas throughout the cloud. In addition, the rates for cooling and line emission can depend on gas density owing to the density dependence of line collisional deexcitation and dielectronic recombination. However, even in this case the ionization parameter remains a convenient means of characterizing the results. The state of the gas is defined by its temperature and by the ion abundances. All ions are predominantly in the ground state, and except for hydrogen and helium the populations of excited levels may be neglected. The relative abundances of the ions of a given element are found by solving the ionization equilibrium equations under the assumption of local balance, subject to the constraint of particle number conservation for each element. Ionization balance is affected by a variety of physical processes, most notably photoionization and radiative and dielectronic recombination. The temperature is found by solving the equation of thermal equilibrium, by equating the net heating of the gas due to absorption of incident radiation with cooling due to emission by the gas. These rates are derived from integrals over the absorbed and emitted radiation spectra. Although Compton scattering is not explicitly included as a source or sink of radiation, its effect is included in the calculation of the thermal balance. The emitted spectrum includes continuum emission by bremsstrahlung and recombination and line emission by a variety of processes including recombination, collisions, and fluorescence following inner shell photoionization. Line transfer is treated using an escape probability formalism and includes the effects of line destruction by collisions and continuum absorption. Transfer of the continuum is calculated using a single stream approximation, as described in Kallman \& McCray (1982). Rates for atomic processes involving electron collisions have been modified since the publication of Kallman \& McCray (1982) to be consistent with those used by Raymond \& Smith (1977). Recombination and ionization rates for Fe have been updated to those of Arnaud \& Raymond (1992). In addition, we have added many optical and UV lines from ions of medium-Z elements (C, N, O, Ne, Si, and S) using collisional and radiative rates from Mendoza (1983). The elements Mg, Ar, Ca, and Ni have also been added. The models have a total of 168 ions, producing 1715 lines, of which 665 have energies greater than 120 eV (10 \AA ), and approximately 800 are resonance lines. For each ion we also calculate the emission from radiative recombination onto all the excited levels which produce resonance lines. The number of such continua is equal to the number of resonance lines in the calculation. Some of the results of the models are sensitive to the rates for dielectronic recombination. We use the high temperature rates given by Aldrovandi \& Pequignot (1973), together with the low temperature rates from Nussbaumer \& Storey (1983). These rates differ significantly for several relevant ions, e.g., Si~III, relative to those of Shull \& Van Steenberg (1983), which were used in most earlier versions of XSTAR. Finally, all models assume element abundances which are close to solar: H:He:C:N:O:Ne:Mg:Si:S:Ar:Ca:Fe:Ni = 1:0.1:3.7E-4:1.1E-4:6.8E-4:2.8E-5:3.5E-5:3.5E-5:1.6E-5:4.5E-06:2.1E-6:2.5E-5:2.0E-6 (\cite{wit71}). An emission line may be produced by two types of physical processes. First is what we will call ``thermal emission,'' which is recombination or collisional excitation by thermal electrons. Second is excitation by photons from the incident continuum. Since this is an elastic scattering process, we will refer to it as ``scattering.'' This process will produce an apparent emission feature if our line of sight to the continuum source is at least partially blocked, and the scattering region must be concentrated on the plane of the sky (see \cite{kro95} for more discussion of these issues). If the scattering region has a bulk velocity greater than the thermal line width then a P~Cygni or inverse P~Cygni profile will form even if the scattering region is spherical. Scattering fails to account for the presence of He~II $\lambda$1640, since this is a subordinate line and the population of the lower level will be negligibly small under conditions appropriate to photoionized gas. Significant opacity in this line would require a temperature greater than $\sim 3\times 10^5$ K and level populations which are close to LTE values. In spite of this, we have investigated the possibility that this mechanism can explain the ratios of the other lines we consider. We assume that the fluxes in scattered lines are proportional to the opacities in the lines. This is likely to be justified if the line optical depths are less than unity, and if the wind ionization balance is approximately uniform. If so, the line flux ratios will equal the opacity ratios. \subsection{Input Parameters} The models we explore fall into two categories: photoionized models and collisional (coronal) models. For each we present the line ratios for both thermal emission and scattering emission mechanisms. The photoionization model parameters are chosen in an attempt to crudely represent the range of observed ionizing spectra from CVs. These typically consist of a soft component which is consistent with a 10--50 eV blackbody, together with a hard X-ray component such as a 10 keV bremsstrahlung (e.g., \cite{cor2a}). The ratio of strengths of these two components varies from one object to another, and with outburst state and subclass. However, a typical ratio is 100:1 (soft:hard) for non-magnetic systems, and 1:1 for magnetic systems (\cite{cor2b}). Beyond such simple considerations, the detailed shape of the ionizing radiation field is difficult to determine accurately. This is due to the strong influence of photoelectric absorption by interstellar and circumstellar gas, and to the limited bandpass and spectral resolution of most past observations in the soft X-ray band (e.g., \cite{ram94}; \cite{mau6a}). Owing to such uncertainties we choose a few very simple ionizing spectra for consideration in our photoionization models: Model A: 30 eV blackbody, Model B: 10 keV brems, Model C: 50 eV blackbody, Model D: 10 eV blackbody. Model G is the mechanically heated (coronal) ionization model. All these models have a total (neutral + ionized) hydrogen column density of $N_{\rm H}=10^{19}~\rm cm^{-2}$, chosen to make them optically thin to the continuum and effectively thin to the escaping resonance lines. In addition, we present two photoionized models which are close to being optically thick, both of which have of column density $N_{\rm H}=10^{23}~\rm cm^{-2}$. Model E has a 30 eV blackbody ionizing spectrum, and Model F has a 10 keV brems ionizing spectrum. For each model we determine the net line flux contained in the wavelength intervals 1237--1243~\AA , 1390--1403~\AA , 1547--1551~\AA , and 1639--1641~\AA , which we refer to as N~V, Si~IV, C~IV, and He~II, respectively. As we will show, these wavelength intervals also contain other lines which can mimic these strong lines, and these other lines are likely to affect the interpretation of the {\it IUE\/} low resolution data as well. The results of our models---the Si~IV/C~IV, N~V/C~IV, and He~II/C~IV line ratios---are summarized in Tables 1--3. For the photoionization models, we consider various values of the ionization parameter for six choices of ionizing continuum shape. For the collisional models, we consider ten values of the gas temperature. We also tested for the dependence on gas density and found that it is negligible for the conditions we consider; all the models presented here have density $n=10^9~\rm cm^{-3}$. \subsection{Model Results} When the line ratios shown in Tables 1--3 are plotted in two diagrams in the same way as the {\it IUE\/} data, several common properties emerge. These are shown in Figures 9 and 10, with various symbols denoting the ionizing spectra: Model A \vrule height6pt width5pt depth-1pt, Model B = $\bullet $, Model C = $\times $, Model D = +, Model E = $\Box $, Model F = $\circ $. We have also tried a model consisting of a 30 eV blackbody together with a 10 keV brems spectrum in a ratio of 99:1; the results are so similar to Model A as to be indistinguishable, i.e., the 30 eV blackbody has far more influence on the ionization balance than does the 10 keV brems in these ratios. Model G, the coronal case, produces line ratio combinations which are almost entirely outside the range spanned by these figures. For the other spectra, as expected, there is little or no dependence on model density for optically thin photoionized models. The shape of the trajectory of the model results in these planes is similar for most of the models. They resemble a {\sf U} shape, although in some cases the right upright of the {\sf U} is missing, and in others it is tipped nearly 45 degrees to the vertical. We can understand the results better if we label the points along the {\sf U} as follows: A = upper left extreme of trajectory; B = bottom of steepest part of left upright; C = lowest point of trajectory; D = bottom of right upright; E = upper right extreme of trajectory. These are shown schematically in Figure~11. In general, the trajectory is traversed from point A to point E as the ionization parameter decreases, and may be understood in terms of the relative ease of ionization of the various ions responsible for line emission. The ionization parameter at which the abundance of a given ion peaks, relative to its parent element, can be derived crudely from the ionization potential. Thus, the four ions responsible for the strongest observed lines may be ordered in terms of decreasing ease of ionization according to: Si~IV, He~II, C~IV, N~V. For the metal ions the ionization parameter at which the emissivities of the lines peak is approximately the same as the ionization parameter at which the ion abundances peak. Thus, at the highest ionization parameters we consider, nitrogen is ionized to or beyond N~V, carbon is ionized beyond C~IV, and silicon is ionized beyond Si~IV. However, the abundance of N~V, and hence its line emissivity, is greater than the corresponding quantities for C~IV, which in turn are greater than for Si~IV. Thus, at high ionization parameter, N~V/C~IV is relatively large and Si~IV/C~IV is small, corresponding to point A of the trajectory in the N~V/C~IV vs.\ Si~IV/C~IV plane. At intermediate ionization parameter, carbon recombines to C~IV and nitrogen recombines below N~V, so that N V/C~IV is small and Si~IV/C~IV is also small (point B). At lower ionization parameter, silicon recombines to Si~IV and carbon may recombine to below C~IV, resulting in an intermediate value of Si~IV/C~IV (point C). At very low ionization parameter, Si~IV/C~IV is at a maximum, and there is an apparent increase in N~V/C~IV (points D, E). This cannot be understood in terms of conventional ionization balance, since the C~IV abundance always exceeds the N~V abundance at low ionization parameter. Rather, it is due to confusion between the N~V 2s--2p doublet at 1238.82, 1242.80~\AA \ and the Mg~II 3s--4p doublet at 1239.93, 1240.39~\AA \ (e.g., \cite{kel87}). Since Mg~II has a lower ionization potential than any of the other ions in question, Mg~II/C~IV increases at low ionization parameter, thus explaining the apparent increase in N~V/C~IV . The behavior of He~II $\lambda$1640 differs from the other lines owing to the fact that it is emitted by recombination, while the others are emitted by collisional excitation. Thus, the emissivity of He~II $\lambda$1640 remains nearly constant at high ionization parameter, while the C~IV $\lambda$1550 line emissivity decreases with increasing ionization parameter. This is in spite of the fact that He~II is more easily ionized than C~IV. This explains the AB part of the trajectory in the He~II/C~IV vs.\ Si~IV/C~IV plane. As the ionization parameter decreases, C~IV recombines to C~III and below, while He~II and He~III are still abundant, thus explaining the DE part of the trajectory. The model behavior under the scattering scenario is qualitatively similar to the thermal excitation scenario, except for the weakness of He~II. The line strengths in the scattering case are less dependent on the gas temperature than in the thermal excitation case, and the oscillator strength for the Mg~II 3s--4p transition is small, so that the trajectory lacks the DE segment in the N~V/C~IV vs.\ Si~IV/C~IV plane. The fact that the coronal models fail almost completely to produce line ratios within the range of our diagrams indicates that it is unlikely that this process dominates in CV line-emitting gas. This result is not surprising, owing to the well-known fact that coronal equilibrium produces ion abundances that have less overlap in parameter space (e.g., temperature) between adjacent ion stages than does photoionization. \subsection{Spectral Dependence} The shape of the ionizing spectrum influences the location of the various points along the trajectory in the two line ratio diagrams, and also the existence of part of the trajectory, most notably the segments between points C and E. For soft spectra, such as the 30 eV blackbody, recombination to species below Si~IV does not occur for the parameter range considered ($\log\xi =-1.5$ to +1.0), so the CDE part of the trajectory is absent in the N~V/C~IV vs.\ Si~IV/C~IV plane. Also, when the ionization parameter is suitable for producing large N~V/C~IV, the Si~IV/C~IV ratio is so small as to be off the scale of Figure~9. For very soft spectra, such as the 10 eV blackbody, there are insufficient hard photons to produce N~V. Thus, points A and B are missing in the N~V/C~IV vs.\ Si~IV/C~IV plane. The DE part of the trajectory is entirely due to Mg~II $\lambda1240$. For hard spectra, such as the 10 keV brems spectrum, X-rays can make Si~IV via Auger ionization even at very low ionization parameter. Thus, the range of Si~IV/C~IV is greatly expanded, and can reach 10 at point D in the N~V/C~IV vs.\ Si~IV/C~IV plane. Other models, such as a 30 eV blackbody, can make large Si~IV/C~IV at low ionization parameter, but they also have very weak N~V line emission, so that the 1240~\AA \ feature is dominated by Mg~II and they are on DE segment of the trajectory. At high ionization parameter He~II is a ``bolometer'' of the ionizing spectrum, since it is dominated by recombination. That is, the He~II strength depends only on the number of photons in the He~II Lyman continuum ($\varepsilon\geq 54.4$ eV). So, harder spectra make stronger He~II, and conversely. This behavior may still hold at point C. At small ionization parameter, He~II is likely to be more abundant than C~IV or Si~IV. This explains the CDE segment of the trajectory in the He~II/C~IV vs.\ Si~IV/C~IV plane. This fact does not seem to depend strongly on spectral shape, although the value of Si~IV/C~IV at points C, D, and E does depend on the spectrum; this ratio increases at all these points for harder spectra, and conversely. The 10 eV blackbody models show a ``hook'' in their trajectory in which the Si~IV/C~IV appears to increase with increasing ionization parameter at the upper end of the range of values we consider. This is counter to the expected behavior of Si~IV at high ionization parameter, since we expect silicon to become ionized beyond Si~IV and the Si~IV abundance to decrease at high ionization parameter. The reason for the model behavior is the confusion between the Si~IV doublet at 1393.76, 1402.77~\AA \ and the O~IV $\rm 2s^22p$--$\rm 2s2p^2$ multiplet at 1397--1407~\AA \ (e.g., \cite{kel87}). The unique behavior of the 10 eV blackbody models is due to the fact that the O~IV lines increase at high ionization parameter only for the softest spectra; harder spectra ionize oxygen past O~IV when other ion abundances are at similar values. Like the N~V and Mg~II lines near 1240~\AA , the Si~IV and O~IV lines near 1400~\AA \ can be confused in {\it IUE\/} low resolution data. The overlap in ionization parameter space of the regions where N~V, Si~IV, C~IV, and He~II predominate does not differ greatly between models with hard X-ray spectra (e.g., 10 keV brems) and those with blackbodies with $kT>30$ eV. A more pronounced difference is due to the fact that the latter spectra have all their ionizing photons crammed into a smaller energy range. Therefore, the ionization parameter scale, which simply counts the energy in ionizing photons, and the distribution of ionization, which really depends on the photon density in the EUV/soft X-ray region, are very different in the two cases. For example, in the 10 keV brems case, C~IV predominates at $\log\xi =0$, while in the 30 eV blackbody case it predominates near $\log\xi=-1$. Thus, our model grid, which spans the range $-3< \log\xi < +1.5$, does not include the region where the gas has recombined below C~IV, etc., in the blackbody case. This accounts for the absence of the CDE part of the line ratio trajectory for the blackbody models. \section{Comparison with Observations} The models presented in the previous section provide a useful context in which to examine the likely physical conditions in CV line-emitting regions. As was discussed earlier, we do not expect these simple models to account for the details of the observed spectra, but we do hope that they will at least crudely reproduce some of the features of the observations. These might include: the range of the observed ratios to within the dispersion of the observed values, i.e., approximately one decade, and possibly the differential behavior of a given source as it varies in time. In fact, we find little evidence for agreement between the observations and any of the models beyond the simplest measures of consistency for some of the line ratios. We begin the comparison of model results with observations by considering the ``normal'' CVs. \subsection{``Normal'' CVs} The observed ratios of ``normal'' CVs lumped together regardless of class (Fig.~6) are clustered within a range of $\sim 1$ decade for log Si~IV/C~IV $\approx -0.5$ and log He~II/C~IV $\approx -1.0$ and $\sim 1.5$ decades for log N~V/C~IV $\approx -0.25$. The larger range of the N~V/C~IV ratio is due largely to the large line ratios of the eclipsing dwarf novae Z~Cha and OY~Car (squares and circles, respectively, in Fig.~6). Otherwise, the range is $\sim 1$ decade centered on $\approx -0.5$. This same general range is spanned by the models, but there is little detailed agreement. One notable failure of the models is the behavior in the He~II/C~IV vs.\ Si~IV/C~IV plane. The photoionization models always produce He~II/C~IV $\gax$ Si~IV/C~IV. This is because, although the He~II line is emitted following radiative recombination and the C~IV and Si~IV lines are formed by collisional excitation, photons at energies greater than 54.4 eV which are responsible for ionizing He~II are also responsible for heating the gas. So, models which efficiently heat the gas and emit Si~IV and C~IV also have efficient ionization of He~II and hence efficient production of the 1640~\AA \ line. In contrast, ``normal'' CVs show He~II/C~IV values which are less than Si~IV/C~IV by $\sim 0.5$ decades. Magnetic systems have He~II/C~IV line ratios which are systematically higher (and Si~IV/C~IV line ratios which are systematically lower) than those of the non-magnetic systems by $\sim 0.25$ decades. Only the DQ~Her stars TV~Col and FO~Aqr (triangles and pentagons, respectively, in Fig.~6) have He~II/C~IV $\geq$ Si~IV/C~IV. It is interesting to note that the ``hook'' in the trajectory of the 10 eV blackbody models referred to in the previous section occurs near the values of these line ratios where most observed objects cluster. Furthermore, the 10 eV models come closest to reproducing the observed He~II/C~IV ratios; they are the only ones for which He~II/C~IV falls below Si~IV/C~IV. The coronal models produce He~II/C~IV values of order 1\% of Si~IV/C~IV when this latter ratio is in the observed range; even at this relatively favorable point, the N~V/C~IV ratio is much smaller than observed. This suggests that the coronal emission mechanism is less likely than photoionization for all CVs. In the N~V/C~IV vs.\ Si~IV/C~IV plane, the models and the data span a similar range of ratios, so there is less indication of the failure of any of the models. The 50 eV blackbody models are least successful at reproducing the most commonly observed values of these ratios simultaneously. The 30 eV blackbody and 10 keV brems models both span the observed range, as do the optically thick models (which also use these ionizing spectra). The ``hook'' in the trajectory of the 10 eV blackbody models causes the Si~IV/C~IV ratio to lie almost entirely in the range $-1.5\leq $ log Si~IV/C~IV $\leq -0.5$, which is close to that spanned by the observed ratios of most ``normal'' CVs. In addition to asking whether there exists a model which can reproduce a given ratio, we can ask whether the distribution of observed ratios is consistent with the analogous model quantity. For example, if the ionizing spectrum and emission mechanism are independent of time, but the luminosity and hence the ionization parameter varies with time, we expect that the ratios of a given object will lie along a {\sf U}-shaped trajectory in the line ratio diagram. In contrast, there appears no clear pattern in the observed ratios for objects with many observations, other than a clustering in a well-defined region of the diagram. \subsection{``Weird'' CVs} Consider next the ``weird'' CVs. Figures~7 and 8 demonstrate that V~Sge, GK~Per, BY~Cam, and AE~Aqr form a sequence of dramatically increasing Si~IV/C~IV and N~V/C~IV at nearly constant He~II/C~IV. In V~Sge and GK~Per, He~II/C~IV $>$ Si~IV/C~IV, unlike most of the ``normal'' CVs, but consistent with the photoionization models. The extreme line ratios of BY~Cam and AE~Aqr are harder to understand. Bonnet-Bidaud \& Mouchet (1987) have suggested a depletion of carbon by a factor of $\sim 10$ to explain the anomalous line ratios of BY~Cam. If a similar deficiency applies to AE~Aqr, depletion by a factor of $\sim 60$ is required. Although such abundance anomalies are possible, they are unlikely to explain the positive correlation between N~V/C~IV and Si~IV/C~IV observed in AE~Aqr and possibly BY~Cam (see Figs.~7 and 8). The observed correlation between these ratios lends support to the hypothesis that confusion between the N~V 2s--2p doublet and the Mg~II 3s--4p doublet, together with a low value of the ionization parameter and hence a large value of the Si~IV/C~IV ratio, is responsible for the apparent anomalous line ratios of BY~Cam and AE~Aqr. However, none of the models reproduce the nearly perfect linear proportionality between N~V/C~IV and Si~IV/C~IV observed in AE~Aqr. There are several possibilities why the models and the observed line ratios are discrepant. First, it is possible that we have failed to consider ionizing spectra of the right type. Since the 10 eV blackbody appears to come closest to providing agreement with the He~II/C~IV and Si~IV/C~IV ratios simultaneously, it is possible that there are confusing lines which we have not included in our models, or which become important under other conditions, which affect the results. Alternatively, the emission region may consist of multiple components with differing physical conditions. If so, the various components must have line ratios which bracket the observed values. This is ruled out by our models: no single set of models brackets the observed ratio of He~II/C~IV, for example. The observed values of this ratio are bracketed by the photoionization models on the high side and the coronal models on the low side, so that a superposition of these models might provide consistent line ratios. However, we consider this possibility to be somewhat contrived, and a more detailed exploration is needed to test whether it can account for all the ratios simultaneously. Another possible explanation for the He~II/C~IV ratio is that our assumption of a stationary steady state is invalid for the line-emitting region. If, for example, the heating and ionization of the gas occurs as the result of many impulsive events, then the time-average value of the ratios could differ significantly from the steady-state model predictions owing to the differing timescales for relaxation of the upper levels of the He~II line from that of C~IV. Such a scenario has been suggested to account for the strength of the He~II lines from the Sun (\cite{ray90}). The differential behavior of the line ratios from a given object could be due at least in part to changes in the ionizing spectrum or ionization mechanism, rather than simply due to changes solely in ionization parameter. This could account for the departures from the variability behavior predicted by the models. In spite of the difficulty in reproducing the observed line ratios, steady-state photoionization models are capable of fitting the absolute strengths of the observed lines (Ko et~al.\ 1996). \section{Summary} We have presented a statistical analysis of the Si~IV/C~IV, N~V/C~IV, and He~II/C~IV emission line ratios of 20 CVs based on $\approx 430$ UV spectra extracted from the {\it IUE\/} ULDA. We find for most systems that these ratios are clustered within a range of $\sim 1$ decade for log Si~IV/C~IV $\approx -0.5$ and log He~II/C~IV $\approx -1.0$ and $\sim 1.5$ decades for log N~V/C~IV $\approx -0.25$. The larger range of the N~V/C~IV ratio is due largely to the large line ratios of the eclipsing dwarf novae Z~Cha and OY~Car; otherwise, the range of log N~V/C~IV is $\sim 1$ decade centered on $\approx -0.5$. The clearest difference between magnetic and non-magnetic CVs is the He~II/C~IV ratio, which is $\sim 0.25$ decades larger in magnetic systems. To place constraints on the excitation mechanism and the physical conditions of the line-emitting gas of CVs, we have investigated the theoretical line ratios of gas in either photoionization and collisional ionization equilibrium. Given the uncertain and variable geometry, density, optical depth of the line-emitting gas and the shape and luminosity of the ionizing spectrum, we have restricted ourselves to consideration of simple slab models each with fixed gas composition, density, and column density. The variables have been the shape and ionization parameter of the ionizing spectrum and the density and column density of the slab; for the collisional models, the temperature was varied. Line emission is produced in these models by recombination or collisional excitation by thermal electrons or by excitation by the ionizing continuum (``scattering''). Within the confines of these simple models, we find little agreement between the observations and any of the models. Specifically, the observed Si~IV/C~IV line ratios are reproduced by many of the models, but the predicted N~V/C~IV line ratios are simultaneously too low by typically $\sim 0.5$ decades. Worse, for no parameters are any of the models able to reproduce the observed He~II/C~IV line ratios; this ratio is far too small in the collisional and scattering models and too large by typically $\sim 0.5$ decades in the photoionization models. Among the latter, the 10 eV blackbody models do the best job of reproducing the three line ratios simultaneously, but the match with the N~V/C~IV line ratio is accomplished only if the observed emission feature near 1240~\AA \ is due to the Mg~II 3s--4p doublet at 1239.93, 1240.39~\AA \ instead of the N~V 2s--2p doublet at 1238.82, 1242.80~\AA . Despite the above generally unfavorable comparisons between observations and simple photoionization and collisional models, our investigation has proven useful in revealing both the problems and promises of understanding the UV line ratios of CVs. Future detailed work could be profitably performed on any and all of the above CV subclasses with more detail in the shape of the ionizing spectrum and the geometrical distribution, density, and column density of the emission region(s). Where the distance is well known, not only line ratios but absolute line strengths can be fit. With larger effective area, weaker lines, less subject to optical depth effects, can be included. With higher spectral resolution, the UV lines be can uniquely identified, thus removing the annoying ambiguity of some of the line identifications. Additional constraints on the physical conditions and optical depth of the line-emitting gas is possible if the UV doublets are resolved. At comparable or slightly higher spectral higher resolution, the velocity field of the line-emitting gas can be constrained to constrain the ionization parameter and hence the density. With realistic photon transport in the models, the line shapes further constrain the models. By extending the bandpass into the far-UV, lines from species with both lower and higher ionization potentials (e.g., C~III, N~III, O~VI, P~V, S~IV, S~VI) provide additional diagnostics. The UV data can and is being obtained with {\it HST\/}, but to obtain the far-UV data, we require the likes of {\it HUT\/} (e.g., \cite{lon96}), {\it ORFEUS\/} (e.g., \cite{ray95}), and {\it FUSE\/}. \acknowledgments We thank John Raymond for useful insights and suggestions and the referee for helpful comments which significantly improved the original manuscript. Work at Lawrence Livermore National Laboratory was performed under the auspices of the U.S.\ Department of Energy under contract No.~W-7405-Eng-48. \clearpage \begin{center} \begin{tabular}{cccc} \multicolumn{4}{c}{\bf TABLE 1}\\ \multicolumn{4}{c}{PHOTOIONIZATION MODEL LINE RATIOS}\\ \tableline \tableline $\log\xi $& log Si~IV/C~IV& log N~V/C~IV& log He~II/C~IV\\ \tableline \multicolumn{4}{l}{Model A: 30 eV Blackbody Spectrum: \vrule height6pt width5pt depth-1pt}\\ \tableline $-1.5$& $?0.00$& $-2.02$& $?0.68$\\ $-1.0$& $-0.55$& $-1.40$& $-0.44$\\ $-0.5$& $-1.12$& $-0.97$& $-0.90$\\ $?0.0$& $-1.56$& $-0.59$& $-0.76$\\ $?0.5$& $-2.20$& $-0.37$& $-0.58$\\ $?1.0$& $-2.92$& $-0.24$& $-0.41$\\ \tableline \multicolumn{4}{l}{Model B: 10 keV Bremsstrahlung Spectrum: $\bullet $}\\ \tableline $-1.5$& $?0.78$& $?0.91$& $?3.65$\\ $-1.0$& $?0.73$& $-0.15$& $?2.99$\\ $-0.5$& $?0.62$& $-1.38$& $?1.87$\\ $?0.0$& $?0.19$& $-1.39$& $?0.52$\\ $?0.5$& $-0.76$& $-1.15$& $-0.64$\\ $?1.0$& $-1.05$& $-0.68$& $-0.84$\\ \tableline \multicolumn{4}{l}{Model C: 50 eV Blackbody Spectrum: $\times $}\\ \tableline $-3.0$& $?1.54$& $?2.18$& $?5.06$\\ $-2.5$& $?1.26$& $?1.04$& $?4.19$\\ $-2.0$& $?0.97$& $-0.25$& $?3.16$\\ $-1.5$& $?0.67$& $-1.66$& $?1.90$\\ $-1.0$& $?0.11$& $-1.69$& $?0.47$\\ $-0.5$& $-1.00$& $-1.17$& $-0.71$\\ $?0.0$& $-2.01$& $-0.97$& $-0.90$\\ $?0.5$& $-2.69$& $-0.62$& $-0.67$\\ $?1.0$& $-3.20$& $-0.26$& $-0.32$\\ $?1.5$& $-3.56$& $?0.22$& $?0.31$\\ \tableline \end{tabular} \end{center} \clearpage \begin{center} \begin{tabular}{cccc} \multicolumn{4}{c}{\bf TABLE 1 --- continued}\\ \multicolumn{4}{c}{PHOTOIONIZATION MODEL LINE RATIOS}\\ \tableline \tableline $\log\xi $& log Si~IV/C~IV& log N~V/C~IV& log He~II/C~IV\\ \tableline \multicolumn{4}{l}{Model D: 10 eV Blackbody Spectrum: +}\\ \tableline $-3.0$& $-1.86$& $?0.35$& $?3.50$\\ $-2.5$& $-1.50$& $-0.90$& $?2.30$\\ $-2.0$& $-0.86$& $-2.23$& $?1.06$\\ $-1.5$& $-0.48$& $-2.90$& $?0.15$\\ $-1.0$& $-0.71$& $-2.52$& $-0.48$\\ $-0.5$& $-0.90$& $-1.88$& $-0.75$\\ $?0.0$& $-0.80$& $-1.08$& $-0.79$\\ $?0.5$& $-0.58$& $-0.31$& $-0.69$\\ $?1.0$& $-0.45$& $?0.25$& $-0.48$\\ $?1.5$& $-0.45$& $?0.57$& $-0.16$\\ \tableline \multicolumn{4}{l}{Model E: 30 eV Blackbody Spectrum, $\log N_{\rm H}=23$: $\Box $}\\ \tableline $-1.5$& $?0.15$& $-1.83$& $?1.46$\\ $-1.0$& $-0.31$& $-1.47$& $?0.18$\\ $-0.5$& $-0.74$& $-1.13$& $-0.49$\\ $?0.0$& $-0.90$& $-0.96$& $-0.59$\\ $?0.5$& $-0.93$& $-0.92$& $-0.58$\\ $?1.0$& $-0.93$& $-0.91$& $-0.58$\\ $?1.5$& $-0.93$& $-0.89$& $-0.58$\\ \tableline \multicolumn{4}{l}{Model F: 10 keV Bremsstrahlung Spectrum, $\log N_{\rm H}=23$: $\circ $}\\ \tableline $-1.5$& $?0.86$& $?1.33$& $?4.20$\\ $-1.0$& $?0.77$& $?0.44$& $?3.49$\\ $-0.5$& $?0.67$& $-0.81$& $?2.42$\\ $?0.0$& $?0.30$& $-1.36$& $?1.05$\\ $?0.5$& $-0.37$& $-1.15$& $-0.07$\\ $?1.0$& $-0.69$& $-0.91$& $-0.44$\\ $?1.5$& $-0.73$& $-0.79$& $-0.48$\\ \tableline \end{tabular} \end{center} \clearpage \begin{center} \begin{tabular}{cccc} \multicolumn{4}{c}{\bf TABLE 2}\\ \multicolumn{4}{c}{SCATTERING MODEL LINE RATIOS}\\ \tableline \tableline $\log\xi $& log Si~IV/C~IV& log N~V/C~IV& log He~II/C~IV\\ \tableline \multicolumn{4}{l}{Model A$^\prime $: 30 eV Blackbody Spectrum: \vrule height6pt width5pt depth-1pt}\\ \tableline $-1.5$& $?0.36$& $-1.44$& $*-9.44$\\ $-1.0$& $-0.26$& $-0.98$& $*-8.84$\\ $-0.5$& $-1.06$& $-0.65$& $*-8.04$\\ $?0.0$& $-1.86$& $-0.33$& $*-6.40$\\ $?0.5$& $-3.09$& $-0.19$& $*-4.86$\\ $?1.0$& $-4.65$& $-0.13$& $*-3.35$\\ $?1.5$& $-6.36$& $-0.10$& $*-2.23$\\ \tableline \multicolumn{4}{l}{Model B$^\prime $: 10 keV Bremsstrahlung Spectrum: $\bullet $}\\ \tableline $-1.5$& $?1.26$& $?0.68$& $-12.02$\\ $-1.0$& $?1.20$& $?0.08$& $-10.71$\\ $-0.5$& $?1.07$& $-0.71$& $*-9.49$\\ $?0.0$& $?0.68$& $-0.86$& $*-8.10$\\ $?0.5$& $-0.59$& $-0.81$& $*-7.14$\\ $?1.0$& $-2.25$& $-0.47$& $*-4.57$\\ $?1.5$& $-5.26$& $-0.06$& $*-0.79$\\ \tableline \multicolumn{4}{l}{Model C$^\prime $: 50 eV Blackbody Spectrum: $\times $}\\ \tableline $-3.0$& $?2.01$& $?1.86$& $-11.36$\\ $-2.5$& $?1.70$& $?1.02$& $-10.81$\\ $-2.0$& $?1.41$& $?0.07$& $-10.32$\\ $-1.5$& $?1.10$& $-0.97$& $*-9.69$\\ $-1.0$& $?0.53$& $-1.24$& $*-8.85$\\ $-0.5$& $-0.75$& $-0.79$& $*-8.45$\\ $?0.0$& $-2.23$& $-0.68$& $*-7.13$\\ $?0.5$& $-4.35$& $-0.42$& $*-5.09$\\ $?1.0$& $-6.98$& $-0.14$& $*-3.19$\\ $?1.5$& $-\infty$& $?0.26$& $*-1.42$\\ \tableline \end{tabular} \end{center} \clearpage \begin{center} \begin{tabular}{cccc} \multicolumn{4}{c}{\bf TABLE 2 --- continued}\\ \multicolumn{4}{c}{SCATTERING MODEL LINE RATIOS}\\ \tableline \tableline $\log\xi $& log Si~IV/C~IV& log N~V/C~IV& log He~II/C~IV\\ \tableline \multicolumn{4}{l}{Model D$^\prime $: 10 eV Blackbody Spectrum: +}\\ \tableline $-3.0$& $-1.28$& $?0.35$& $-11.04$\\ $-2.5$& $-1.24$& $-0.48$& $-10.32$\\ $-2.0$& $-0.61$& $-1.41$& $*-9.50$\\ $-1.5$& $-0.13$& $-2.22$& $*-9.62$\\ $-1.0$& $-0.45$& $-2.06$& $*-9.77$\\ $-0.5$& $-0.85$& $-1.48$& $*-9.57$\\ $?0.0$& $-1.08$& $-0.74$& $*-8.62$\\ $?0.5$& $-1.16$& $-0.02$& $*-7.52$\\ $?1.0$& $-1.19$& $?0.51$& $*-6.85$\\ $?1.5$& $-1.20$& $?0.81$& $*-6.52$\\ \tableline \multicolumn{4}{l}{Model E$^\prime $: 30 eV Blackbody Spectrum, $\log N_{\rm H}=23$: $\Box $}\\ \tableline $-1.5$& $?2.01$& $?0.92$& $*-6.19$\\ $-1.0$& $?1.85$& $?0.27$& $*-5.06$\\ $-0.5$& $?1.85$& $-0.03$& $*-3.90$\\ $?0.0$& $?1.90$& $-0.01$& $*-2.24$\\ $?0.5$& $?1.90$& $-0.02$& $*-0.77$\\ $?1.0$& $?1.92$& $-0.18$& $*?0.58$\\ $?1.5$& $?1.96$& $?0.09$& $*?1.91$\\ \tableline \multicolumn{4}{l}{Model F$^\prime $: 10 keV Bremsstrahlung Spectrum, $\log N_{\rm H}=23$: $\circ $}\\ \tableline $-1.5$& $?1.23$& $?1.21$& $-11.89$\\ $-1.0$& $?1.19$& $?0.77$& $-10.29$\\ $-0.5$& $?1.19$& $?0.28$& $*-8.47$\\ $?0.0$& $?1.19$& $-0.21$& $*-6.36$\\ $?0.5$& $?1.14$& $-0.62$& $*-4.50$\\ $?1.0$& $?1.08$& $-0.79$& $*-1.93$\\ $?1.5$& $?1.05$& $-0.81$& $*?1.19$\\ \tableline \end{tabular} \end{center} \clearpage \begin{center} \begin{tabular}{cccc} \multicolumn{4}{c}{\bf TABLE 3}\\ \multicolumn{4}{c}{COLLISIONAL IONIZATION MODEL LINE RATIOS}\\ \tableline \tableline $T$ (10{,}000 K)& log Si~IV/C~IV& log N~V/C~IV& log He~II/C~IV\\ \tableline \multicolumn{4}{l}{Model G: Constant Temperature}\\ \tableline *3& $-0.12$& $-2.63$& $-1.67$\\ *5& $-0.46$& $-3.00$& $-2.61$\\ *7& $-1.83$& $-2.91$& $-3.54$\\ *9& $-2.65$& $-2.79$& $-4.04$\\ 11& $-2.80$& $-2.16$& $-3.99$\\ 13& $-2.65$& $-1.29$& $-3.58$\\ 15& $-2.44$& $-0.52$& $-3.13$\\ 17& $-2.29$& $?0.04$& $-2.74$\\ 19& $-2.17$& $?0.38$& $-2.42$\\ 21& $-2.09$& $?0.52$& $-2.16$\\ \tableline \end{tabular} \end{center} \clearpage
proofpile-arXiv_065-436
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} The Mott transition is one of the most fascinating phenomenons arising from electron-electron interactions, and occurs in a wide range of materials \cite{mott_metal_insulator}. In fact two different Mott transitions exist: one can either stay at a given (commensurate) filling and vary the strength of the interactions (I call this transition Mott-U and it occurs in e.g. vanadium oxides or in organic (quasi-)one dimensional systems), one can also keep the strength of the interaction constant and dope the system to move away from the commensurate density (a Mott-$\delta$ transition, a situation realized in High Tc superconductors or in quantum wires). Although the basic underlying physics behind these transitions is by now well understood it has proved incredibly difficult to tackle it in either $d=2$ or $d=3$ due to our lack of tools to treat strongly interacting systems \cite{mott_mean_fields}. In fact nearly all the fine points of the transition, such as the critical properties or the transport properties remain unknown. One dimension constitute a special case where a rather complete study of the Mott transition can be done. This offers special interest both for theoretical and experimental reasons. From a theoretical point of view, the effect of e-e interactions is particularly strong and leads to a non-fermi liquid state (the so called Luttinger liquid (LL)). One can therefore expect drastic effects on the transport properties of the system. From the experimental point of view, both transitions at constant doping and by varying the doping can be realized, e.g. in organic conductors \cite{jerome_revue_1d} and quantum wires \cite{tarucha_quant_cond} or Josephson junction networks \cite{vanoudenaarden_josephson_mott}. Although the thermodynamic properties of the Mott transition were understood a long time ago for the Hubbard model, which was shown to be a Mott insulator at half filling \cite{lieb_hubbard_exact,emery_revue_1d,solyom_revue_1d}, very little was known of the Mott-$\delta$ transition and of the transport properties: a parquet treatment gave the effective scattering \cite{gorkov_pinning_parquet} but was limited to half filling and small (perturbative) interactions, and only the zero frequency conductivity (i.e. the Drude weight) could be computed by Bethe-Ansatz for the particular case of the Hubbard model \cite{shastry_twist_1d,schulz_conductivite_1d}). Recently a complete picture of both Mott transitions as well as a full description of the transport properties $\sigma(\omega,T,\delta)$ for any commensurate filling, both for bosons or fermions, was obtained \cite{giamarchi_umklapp_1d,giamarchi_attract_1d,giamarchi_curvature}. In these proceedings I will review some of these results. No derivation will be given and the reader is referred to \cite{giamarchi_umklapp_1d,giamarchi_attract_1d,giamarchi_curvature} for derivation as well as the complete analytical expressions, for the figures presented here. Such a presentation is done in section~\ref{section2} where umklapp effects are presented and in section~\ref{section3} where both the critical properties of the Mott transition(s) and the transport properties are examined. In section~\ref{section4} application of these results to the physics of the organic materials is done. \section{Lattice effects and umklapp terms} \label{section2} In the continuum e-e interactions conserve momentum, and thus current, and cannot lead to any finite conductivity as a consequence of Galilean invariance. In the presence of a lattice however the momentum need only to be conserved modulo one vector of the reciprocal lattice, and such interaction process (named umklapp process) can lead to finite resistivity. In a fermi liquid, umklapps are responsible for the intrinsic resistivity $\rho(T)\sim T^2$. In one dimension it was rapidly realized \cite{emery_revue_1d,solyom_revue_1d} that umklapps are also responsible for the Mott-U transition at half filling. Away from half filling they are ``frozen'' due to the mismatch in momentum, and are usually discarded as irrelevant: the system becomes then a perfect metal. However both for the transport properties and to study the Mott-$\delta$ transition it is necessary to have a description of the umklapp processes even for finite doping. This can be achieved using the so called bosonization representation, that uses that {\bf all} excitations of a one dimensional system can be described in term of density oscillations \cite{emery_revue_1d,solyom_revue_1d,haldane_bosonisation}. The charge properties of a {\bf full} interacting one dimensional system (excluding umklapp terms) is therefore described by \begin{equation} \label{quadra} H_0 = \frac1{2\pi} \int dx \; u_\rho K_\rho (\pi\Pi_\rho)^2 + \frac{u_\rho}{K_\rho} (\nabla \phi)^2 \end{equation} where $\nabla \phi = \rho(x)$, $\rho(x)$ is the charge density and $\Pi$ is the conjugate momentum to $\phi$. All the interaction effects are hidden in the parameters $u_\rho$ (the velocity of charge excitations) and $K_\rho$ (the Luttinger liquid exponent controlling the decay of all correlation functions). This description (\ref{quadra}) is valid for an arbitrary one-dimensional interacting system, {\bf provided} one uses the proper $u$ and $K$ (in the following I will drop the $\rho$ index). In a general way $K =1$ is the noninteracting point, $K>1$ means attraction whereas $K<1$ means repulsion. The umklapp process can also be given in terms of boson operators \cite{emery_revue_1d,solyom_revue_1d}. In fact umklapps exist not only at half filling but for higher commensurabilities as well by transferring more particles across the fermi surface (such processes are generated in higher order in perturbation theory) \cite{giamarchi_curvature,schulz_mott_revue}. For even commensurabilities the Hamiltonian corresponding to the umklapp process is \begin{equation} \label{um1n} H_{\frac1{2n}} = g_{\frac1{2n}} \int dx \cos(n \sqrt8 \phi_\rho(x) + \delta x) \end{equation} where $n$ is the order of the commensurability ($n=1$ for half filling - one particle per site; $n=2$ for quarter filling - one particle every two sites and so on). The coupling constant $g_{1/2n}$ is the umklapp process corresponding to the commensurability $n$ and $\delta$ the deviation (doping) from the commensurate filling. $g_{\frac12}$ corresponds to one particle per site (half filling) For simple models such as the Hubbard model $g_{\frac12}=U$, but this does not need to be the case for more general models (see in particular section~\ref{section4}). For $1/4$ filling and a typical interaction $U$ one has $g_{1/4}\sim U (U/E_F)^3$. Odd commensurability involves spin \cite{schulz_mott_revue} but can be treated similarly. Similar expressions can be derived for the case of bosons \cite{giamarchi_attract_1d}. It is therefore remarkable that in one dimension $H_0+H_{\frac1{2n}}$ provides the solution to {\bf all} Mott transitions, for {\bf all} systems and {\bf all} (for particles with spin: even) commensurabilities. \section{Mott transition(s)} \label{section3} Let us now examine the physical properties of the Mott transitions close to a commensurability of order $n$ described by $H_0+H_{\frac1{2n}}$. \subsection{Phase diagram} The Mott-U and Mott-$\delta$ transitions are radically different and lead to the phase diagram shown in Figure~\ref{phasediag}. \begin{figure} \centerline{\epsfig{file=figure1.eps,angle=-90,width=7cm}} \caption{Phase diagram close to a commensurability of order $n$ ($n=1$ for half filling and $n=2$ for quarter filling). $U$ denotes a general (i.e. not necessarily local) repulsion. $\mu$ is the chemical potential and $\delta$ the doping. MI means Mott insulator and LL Luttinger liquid (metallic) phase. The critical exponent $K_c$ and velocity $u$ depends on whether it is a Mott-U or Mott-$\delta$ transition.} \label{phasediag} \end{figure} The Mott-U is of the Kosterlitz Thouless type \cite{emery_revue_1d,solyom_revue_1d} and occurs for a critical value of $K$, $K_c=1/n^2$ \cite{giamarchi_curvature,schulz_mott_revue}. For half filling the transition point is the noninteracting one ($K_c=1$) but for higher commensurability one reaches the Mott insulator only for very repulsive interactions (for example for quarter filling $K_c^U=1/4$). In the metallic phase the system is a LL, with finite compressibility and Drude weight. The Mott insulator has a gap in the charge excitations (thus zero compressibility). At the transition there is a finite jump both in the compressibility and Drude weight. The dynamical exponent is $z=1$. To study the Mott-$\delta$ transition it is useful \cite{giamarchi_umklapp_1d} to map the sine-gordon Hamiltonian $H_0+H_{\frac1{2n}}$ to a spinless fermion model (known as massive Thiring model \cite{emery_revue_1d,solyom_revue_1d}), describing the charge excitations (solitons) of the sine-gordon model. The remarkable fact is that {\bf close} to the Mott-$\delta$ transition the solitons become non-interacting, and one is simply led to a simple semi-conductor picture of two bands separated by a gap (see figure~\ref{Thiring}). \begin{figure} \centerline{\epsfig{file=figure2.eps,angle=-90,width=6cm}} \caption{Lower Hubbard band and Upper Hubbard band. This concept can be made rigorous in one dimension by mapping the full interacting system to a massive Thiring model. Optical transitions can be made either within or between the two ``bands''.} \label{Thiring} \end{figure} This image has to be used with caution since the solitons are only non-interacting for infinitesimal doping (or for a very special value of the initial interaction) and has to be supplemented by other techniques \cite{giamarchi_umklapp_1d}. Nevertheless it provides a very appealing description of the LHB and UHB and a good guide to understand the phase diagram and transport properties. The Mott-$\delta$ transition is of the commensurate-incommensurate type. The {\bf universal} (independent of the interactions) value of the exponents $K_c^\delta = 1/(2n^2)$ is half of the one of Mott-U transition. Since at the Mott-$\delta$ transition the chemical potential is at the bottom of a band the velocity goes to zero with doping. This leads to a continuous vanishing of the Drude weight and compressibility. The dynamical exponent is now $z=2$. For more details see \cite{giamarchi_umklapp_1d,giamarchi_curvature,schulz_mott_revue,mori_mott_1d}. For bosons, the phase diagram of figure~\ref{phasediag} is well compatible with numerical results \cite{batrouni_bosons_numerique} and higher dimensional proposals \cite{fisher_boson_loc}. \subsection{Transport properties} Let us now look at the transport properties. The full conductivity (real and imaginary part) $\sigma(\omega,T,\delta)$ can be found in \cite{giamarchi_umklapp_1d,giamarchi_attract_1d,giamarchi_curvature} and we just examine here simple limits. The ac conductivity (at $T=0$) for $\delta=0$ is shown in figure~\ref{sigacdel0}. In the Mott insulator $\sigma$ is zero until $\omega$ can make transitions between the LHB and UHB. At the threshold one has the standard square root singularity coming from the density of states (see figure~\ref{Thiring}). For higher frequencies interactions dress the umklapps and give a nonuniversal (i.e. interaction-dependent) power law-like decay. Such a power law is beyond the reach of the simple noninteracting description of Figure~\ref{Thiring}. \begin{figure} \centerline{\epsfig{file=figure3.eps,angle=-90,width=7cm}} \caption{ac conductivity for $\delta=0$ for a commensurability of order $n$. $\Delta$ is the Mott gap. The full line is the conductivity in the Mott insulator. The dashed one is $\sigma$ in the metallic regime. It contains both a Drude peak of weight $D$ and a regular part.} \label{sigacdel0} \end{figure} Away from commensurate filling ($\delta\ne 0$) the conductivity is shown in Figure~\ref{sigacdeln0} (only the case where the half filled system is a MI is shown. For the other case see \cite{giamarchi_attract_1d}). \begin{figure} \centerline{\epsfig{file=figure4.eps,angle=-90,width=7cm}} \caption{ac conductivity for $\delta\ne 0$ for a commensurability of order $n$. $\Delta$ is the Mott gap. In addition to the Drude peak of weight $D$ the regular part has two distinct regimes. } \label{sigacdeln0} \end{figure} Features above the Mott gap are unchanged (the system has no way to know it is or not at half filling at high frequencies). The two new features are a Drude peak with a weight proportional to $\delta/\Delta$, and an $\omega^3$ absorption \cite{giamarchi_curvature} at small frequency. Features above the Mott gap come from inter (hubbard)-band transitions whereas they come from intra-band processes below the Mott gap (see figure~\ref{Thiring}). The dc conductivity can be computed by the same methods and is shown in figure~\ref{dcsig}. \begin{figure} \centerline{\epsfig{file=figure5.eps,angle=-90,width=7cm}} \caption{dc conductivity as a function of $T$. Full line is for the Mott insulator, dashed line is in the metallic regime. $\Delta$ is the Mott gap.} \label{dcsig} \end{figure} Here again the dressing of umklapps by the other interactions results in a nonuniversal power law dependence. If the interactions are repulsive enough the resistivity can even {\bf increase} as a function of temperature well above the Mott gap. Two universal behavior are expected: at the Mott-U transition one has $\rho(T)\sim T$ and $\sigma(\omega)\sim 1/(\omega\ln(\omega)^2)$, whereas at the Mott-$\delta$ transition due to the different $K_c$ one expects $\rho(T)\sim 1/T$. All this results are completely general and apply to any one-dimensional systems for which $\Delta$ is smaller than the scale above which {\bf all} interactions can be treated perturbatively (typically $U$), a situation that covers most of the experimentally relevant cases (see section~\ref{section4}). It is noteworthy that the above results are also valid in the presence not of umklapp processes but of a simple periodic potential (the lattice corresponds itself to a $4k_F$ periodic potential). For a $2k_F$ periodic potential transport properties are similar to the one above with the replacement of $4n^2 K_\rho$ by $1+K_\rho$. \section{Organic compounds} \label{section4} The above results have a direct application to organic conductors. These compounds are $1/4$ filled by chemistry but due to a slight dimerization of the chain an half filled umklapp $g_{1/2}\sim U (D/E_F)$ also exists where $D$ is the dimerization gap, and $U$ a typical strength of the interactions \cite{jerome_revue_1d}. Since $D/E_F$ is quite small the umklapp term is much smaller than the other interactions leading to a quite small Mott gap (see e.g. \cite{penc_numerics} for a numerical estimation of the parameters). There is also a $1/4$ filled umklapp $g_{1/4}\sim U (U/E_F)^3$, which is as we saw less relevant but can be depending on the typical interaction $U$ much larger in magnitude than $g_{1/2}$. Since the organic conductors are only quasi-one dimensional systems with a perpendicular hopping integral $t_\perp$ between the chains one can distinguish various domains in energy scale (temperature or frequency) as shown in figure~\ref{scales} \begin{figure} \centerline{\epsfig{file=figure6.eps,angle=-90,width=7cm}} \caption{Four important energy regimes for quasi-one dimensional systems. In ``pert.'' everything can be treated perturbatively. In ``1D'' the interactions lead to the one dimensional physics, and hopping from chain to chain is incoherent. In ``2D-3D'' the hoping between chains is coherent. The system orders in ``Ordered''.} \label{scales} \end{figure} The most relevant questions being of course: what is the strength of the interactions in these systems, what is the scale for $T_{\rm cr}$ (the bare $t_\perp$ or lower \cite{hopping_general}), and what is the physics below $T_{\rm cr}$. In the absence of $t_\perp$ one expects therefore these compounds to be Mott insulators. This is the case for the TMTTF family that has indeed a conductivity \cite{creuzet_tmttf} similar to the one of figure~\ref{dcsig} (full line). Indeed if $\Delta > T_{\rm cr}$, one expects the Mott gap to render the single particle hopping $t_\perp$ irrelevant ($E_{\rm cr}$ would thus not exist). This family should be described by one-dimensional physics. Further check of this can be provided by examination of the optical (ac) conductivity, and comparing it to figure~\ref{sigacdel0}. Measurements of the transverse conductivity would also give information on the relevance of the transverse hopping. Note that the temperature dependence of the dc resistivity and the frequency dependence of the optical conductivity provide a {\bf direct} measure of the $K_\rho$ exponent of the Luttinger Liquid and give therefore crucial information on the importance of interactions in such systems (the optical conductivity has the advantage to be free from thermal expansion problems). A naive fit in TMTTF would give a value of $K_\rho\sim 0.8$, widely different from the one of $K_\rho=0.3$ extracted from the NMR \cite{wzietek_tmtsf_nmr}. A way to get out of this predicament could be that the conductivity is in fact dominated by $1/4$ filling umklapp processes till very close to the Mott gap giving \begin{equation} \rho(T) \sim g_{1/2}^2 T^{4K-3} + g_{1/4}^2 T^{16K-3} \sim g_{1/4}^2 T^{16K-3} \end{equation} but this point clearly deserves further investigation. On the other hand, the TMTSF family shows a rather good metallic behavior with a $T^2$ resistivity, indicating the importance of transverse hopping. This is to be expected if $\Delta > E_{\rm cr}$. There is important controversy on the value of $E_{\rm cr}$ \cite{behnia_transport_magnetic,gorkov_sdw_tmtsf}. Regardless of the value of $E_{\rm cr}$ the physics for $(\omega, T ) > E_{\rm cr}$ will still be controlled by one dimensional effects. Indeed For the TMTSF family the optical conductivity \cite{dressel_optical_tmtsf} is very well compatible with the figure~\ref{sigacdeln0}. In particular the optical peak can easily be interpreted in term of the Mott insulator described here. Of course more detailed comparison of the structure above the gap an in particular a check for the power law decay of figure~\ref{sigacdeln0} would be worthy to do. The low energy features and in particular the metallic behavior are closer to the {\bf doped} system rather than the commensurate one. ``doping'' is not too surprising since if one particle hopping between chains is relevant, one expect small deviations to the commensurate filling due to the warping of the Fermi surface. One therefore expect a very small spectral weight in the $\delta(\omega)$ part. Since one has a clear idea of the (purely) one-dimensional conductivity (figure~\ref{sigacdeln0}), a detailed comparison with experimental data should provide an indication on the value of $E_{\rm cr}$. The question on whether the physics below $E_{\rm cr}$ is simply ``fermi liquid'' like \cite{gorkov_sdw_tmtsf} or still retains some features of one-dimensionality and interactions is still open. One way to settle this issue is a detailed examination of the low frequency part of the optical conductivity and measurements of the transverse dc conductivity in this regime. Another way would be to examine the effects of impurities on the dc conductivity. Indeed one expect drastic localization in a one dimensional regime and very weak effects for a FL \cite{giamarchi_loc}. {\bf Acknowledgments:} It is a pleasure to thank L. Degiorgi, L.P. Gor'kov, G. Gr\"uner, D. J\'erome, A.J. Millis, H.J. Schulz and B.S. Shastry for many interesting discussions.
proofpile-arXiv_065-437
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} \vskip 10pt The precise measurement of the $W$-boson mass $M_W$ constitutes a primary task of the forthcoming experiments at the high energy electron--positron collider LEP2 ($2 M_W \leq \sqrt{s} \leq 210$~GeV). A meaningful comparison between theory and experiment requires an accurate description of the fully exclusive processes $e^+ e^- \to 4f$, including the main effects of radiative corrections, with the final goal of providing predictions for the distributions measured by the experiments. A large effort in the direction of developing tools dedicated to the investigation of this item has been spent within the Workshop ``Physics at LEP2'', held at CERN during 1995. Such an effort has led to the development of several independent four-fermion codes, both semianalytical and Monte Carlo, extensively documented in~\cite{wweg}. {\tt WWGENPV} is one of these codes, and the aim of the present paper is to describe in some detail the developments performed with respect to the original version~\cite{cpcww}, where a description of the formalism adopted and the physical ideas behind it can be found. As discussed in~\cite{wmass}, the most promising methods for measuring the $W$-boson mass at LEP2 are the so called ``threshold'' and ``direct reconstruction'' methods. For the first one, a precise evaluation of the threshold cross section is required. For the second one, a precise description of the invariant-mass shape of the hadronic system in semileptonic and hadronic decays is mandatory. In order to meet these requirements, the previous version of the program has been improved, both from the technical and physical point of view. On the technical side, in addition to the ``weighted event integration'' and ``unweighted event generation'' branches, the present version can also be run as an ``adaptive Monte Carlo'' integrator, in order to obtain high numerical precision results for cross sections and other relevant observables. In the ``weighted event integration'' branch, a ``canonical'' output can be selected, in which several observables are processed in parallel together with their most relevant moments~\cite{wweg}. Moreover, the program offers the possibility of generating events according to a specific flavour quantum number assignment for the final-state fermions, or of generating ``mixed samples'', namely a fully leptonic, fully hadronic or semileptonic sample. On the physical side, the class of tree-level EW diagrams taken into account has been extended to include all the single resonant diagrams ({\tt CC11/CC20}), in such a way that all the charged current processes are covered. Motivated by the physical relevance of keeping under control the effects of the transverse degrees of freedom of photonic radiation, both for the $W$ mass measurement and for the detection of anomalous couplings, the contribution of QED radiation has been fully developed in the leading logarithmic approximation, going beyond the initial-state, strictly collinear approximation, to include $p_T / p_L $ effects both for initial- and final-state photons. Last, an hadronic interface to {\tt JETSET} in the generation branch has been added. In the present version the neutral current backgrounds are neglected in the fully hadronic and leptonic decay channels, but this is not a severe limitation of the program since, at least in the LEP2 energy range, these backgrounds can be suppressed by means of proper invariant mass cuts. On the other hand, the semileptonic decay channels are complete at the level of the Born approximation EW diagrams ({\tt CC11/CC20} diagrams), and this feature allows to treat at best those channels that are expected to be the most promising for the direct reconstruction of the $W$ mass, because free of systematics such as the ``color reconnection'' and the ``Bose-Einstein correlation'' problems. In the development of the code, particular attention has been paid to the possibility of obtaining precise results in relatively short CPU time. As shown in~\cite{wweg}, {\tt WWGENPV} is one of the most precise four fermion Monte Carlo's from the numerical point of view. This feature allows the use of the code also for fitting purposes. \noindent \section{The most important new features} \vskip 10pt In the following we list and briefly describe the most important technical and physical developments implemented in the new version of {\tt WWGENPV}. \vskip 24pt \noindent TECHNICAL IMPROVEMENTS The present version of the program consists of three branches, two of them already present in the original version but upgraded in some respect, the third one completely new. \noindent \begin{itemize} \item Unweighted event generation branch. This branch, meant for simulation purposes, has been improved by supplying an option for an hadronisation interface (see more details later on). \item Weighted event integration branch. This branch, intended for computation only, includes as a new feature an option for selecting a ``canonical'' output containing predictions for several observables and their most relevant moments together with a Monte Carlo estimate of the errors. According to the strategy adopted in~\cite{wweg}, the first four Chebyshev/power moments of the following quantities are computed: the production angle of the $W^+$ with respect to the positron beam ({\tt TNCTHW, N=1,2,3,4}), the production angle $\vartheta_{d}$ of the down fermion with respect to the positron beam ({\tt TNCTHD}), the decay angle $\vartheta_{d}^*$ of the down fermion with respect to the direction of the decaying $W^-$ measured in its rest frame ({\tt TNCTHSR}), the energy $E_{d}$ of the down fermion normalized to the beam energy ({\tt XDN}), the sum of the energies of all radiated photons ({\tt XGN}) normalized to the beam energy, the lost and visible photon energies normalized to the beam energy ({\tt XGNL} and {\tt XGNV}), respectively, and, finally \begin{eqnarray*} <x_m> \, = \, {1 \over \sigma} \, \int \left( { { \sqrt{s_+} + \sqrt{s_-} -2 M_W} \over {2 E_b} } \right) \, d \sigma \end{eqnarray*} where $s_+$ and $s_-$ are the invariant masses of the $W^+$ and $W^-$ decay products, respectively ({\tt XMN}). \item Adaptive integration branch. This new branch is intended for computation only, but offers high precision performances. On top of the importance sampling, an adaptive Monte Carlo integration algorithm is used. The code returns the value of the cross section together with a Monte Carlo estimate of the error. Moreover, if QED corrections are taken into account, also the average energy and invariant mass losses are printed. The program must be linked to NAG library for the Monte Carlo adaptive routine. Full consistency between non-adaptive and adaptive integrations has been explicitly proven. \end{itemize} In each of these three branches the user is asked to specify the four-fermion final state which is required. The final states at present available are those containing {\tt CC03} diagrams as a subset. Their list, as appears when running the code, is the following: \begin{verbatim} PURELY LEPTONIC PROCESSES [0] ---> E+ NU_E E- BAR NU_E [1] ---> E+ NU_E MU- BAR NU_MU [2] ---> E- BAR NU_E MU+ NU_MU [3] ---> E+ NU_E TAU- BAR NU_TAU [4] ---> E- BAR NU_E TAU+ NU_TAU [5] ---> MU+ NU_MU MU- BAR NU_MU [6] ---> MU+ NU_MU TAU- BAR NU_TAU [7] ---> MU- BAR NU_MU TAU+ NU_TAU [8] ---> TAU+ NU_TAU TAU- BAR NU_TAU SEMILEPTONIC PROCESSES [9] ---> E+ NU_E D BAR U [10] ---> E- BAR NU_E BAR D U [11] ---> E+ NU_E S BAR C [12] ---> E- BAR NU_E BAR S C [13] ---> MU+ NU_MU D BAR U [14] ---> MU- BAR NU_MU BAR D U [15] ---> MU+ NU_MU S BAR C [16] ---> MU- BAR NU_MU BAR S C [17] ---> TAU+ NU_TAU D BAR U [18] ---> TAU- BAR NU_TAU BAR D U [19] ---> TAU+ NU_TAU S BAR C [20] ---> TAU- BAR NU_TAU BAR S C HADRONIC PROCESSES [21] ---> D BAR U BAR D U [22] ---> D BAR U BAR S C [23] ---> S BAR C BAR D U [24] ---> S BAR C BAR S C MIXED SAMPLES [25] ---> LEPTONIC SAMPLE [26] ---> SEMILEPTONIC SAMPLE [27] ---> HADRONIC SAMPLE \end{verbatim} It is worth noting that in the weighted event integration and unweighted event generation branches, besides the possibility of selecting a specific four-fermion final state, an option is present for considering three realistic mixed samples corresponding to the fully leptonic, fully hadronic or semileptonic decay channel, respectively. When the generation of a mixed sample is required, as a a first step the cross section for each contributing channel is calculated; as a second step, the unweighted events are generated for each contributing channel with a frequency given by the weight of that particular channel with respect to the total. In the generation branch, if the hadronisation interface is not enabled, an $n$-tuple is created with the following structure: \begin{verbatim} 'X_1','X_2','EB' ! x_{1,2} represent the ! energy fractions of incoming e^- and ! e^+ after ISR; EB is the beam energy; 'Q1X','Q1Y','Q1Z','Q1LB' ! x,y,z components of the momentum ! of particle 1 and the particle label ! according to PDG; the final-state ! fermions are assumed to be massless; 'Q2X','Q2Y','Q2Z','Q2LB' ! as above, particle 2 'Q3X','Q3Y','Q3Z','Q3LB' ! " " " 3 'Q4X','Q4Y','Q4Z','Q4LB' ! " " " 4 'AK1X','AK1Y','AK1Z' ! x,y,z components of the momentum of ! the photon from particle 1; ! they are 0 if no FSR has been chosen ! and/or if particle 1 is a neutrino; 'AK2X','AK2Y','AK2Z' ! as above, particle 2 'AK3X','AK3Y','AK3Z' ! " " " 3 'AK4X','AK4Y','AK4Z' ! " " " 4 'AKEX','AKEY','AKEZ' ! x,y,z components of the momentum of ! the photon from the initial-state ! electron; they are 0 if no ISR has ! been chosen; 'AKPX','AKPY','AKPZ' ! as above, initial-state positron; \end{verbatim} If the hadronisation interface is enabled, fully hadronised events are instead made available to a user routine in the {\tt /HEPEVT/} format (see below). \vskip 24pt \noindent PHYSICAL IMPROVEMENTS \\ The main theoretical developments with respect to the original version concern the inclusion of additional matrix elements to the tree-level kernel and a more sophisticated treatment of the photonic radiation, beyond the initial-state, strictly collinear approximation. Moreover, an hadronisation interface to { \tt JETSET} has been also provided. {\it Tree-level EW four-fermion diagrams} -- In addition to double-resonant charged-current diagrams {\tt CC03} already present in the previous version, the matrix element includes also the single-resonant charged-current diagrams {\tt CC11} for $\mu $ and $\tau$'s in the final state, and {\tt CC20} for final states containing electrons. This allows a complete treatment at the level of four-fermion EW diagrams of the semileptonic sample, which appears the most promising and cleanest for the direct mass reconstruction method due to the absence of potentially large ``interconnection'' effects~\cite{wmass}. Concerning {\tt CC20} diagrams, the importance sampling technique has been extended to take care of the peaking behaviour of the matrix element when small momenta of the virtual photon are involved. As a consequence of the fact that the tree-level matrix-element is computed in the massless limit, a cut on the minimum electron (positron) scattering angle must be imposed. The inclusion of such a cut eliminates the problems connected with gauge-invariance in the case of {\tt CC20} processes, for which the present version does not include any so-called ``reparation'' scheme~\cite{bhf,wwcd}. Anyway, when for instance a set of ``canonical cuts''~\cite{wweg} is used, the numerical relevance of such gauge-invariance restoring schemes has been shown~\cite{bhf,wwcd} to be negligible compared with the expected experimental accuracy. {\it Photonic corrections} -- As far as photonic effects are concerned, the original version, as stated above, included only leading logarithmic initial-state corrections in the collinear approximation within the SF formalism. The treatment has been extended in a two-fold way: the contribution of final-state radiation has been included and the $p_T / p_L$ effects have been implemented both for initial- and final-state radiation. The inclusion of the transverse degrees of freedom has been achieved by generating the fractional energy $x_{\gamma}$ of the radiated photons by means of resummed electron structure functions $D(x; s)$~\cite{sf} ($x_{\gamma} = 1 - x$) and the angles using an angular factor inspired by the pole behaviour $1 / (p \cdot k)$ for each charged emitting fermion. This allows to incorporate leading QED radiative corrections originating from infrared and collinear singularities, taking into account at the same time the dominant kinematic effects due to non-strictly collinear photon emission, in such a way that the universal factorized photonic spectrum is recovered. According to this procedure, the leading logarithmic corrections from initial- and final-state radiation are isolated as a gauge-invariant subset of the full calculation (not yet available) of the electromagnetic corrections to $e^+ e^- \to 4f$. Due to the inclusion of $p_T$-carrying photons at the level of initial-state radiation, the Lorentz boost allowing the reconstruction of the hard-scattering event from the c.m. system to the laboratory one has been generalized to keep under control the $p_T$ effects on the beam particles. Final state radiation and $p_T / p_L$ effects are not taken into account in the adaptive integration branch. {\it Non-QED corrections} -- Coulomb correction is treated as in the original version on double-resonant {\tt CC03} diagrams. QCD corrections are implemented in the present version in the naive form according to the recipe described in~\cite{wweg,wwcd}. The treatment of the leading EW contributions is unchanged with respect to the original version. {\it Hadronisation} -- Final-state quarks issuing from the electroweak 4-fermion scattering are not experimentally observable. An hadronisation interface is provided to the {\tt JETSET} package~\cite{jetset} to allow events to be extrapolated to the hadron level, for example for input to a detector simulation program. Specifically, the 4-fermion event structure is converted to the {\tt /HEPEVT/} convention, then {\tt JETSET} is called to simulate QCD partonic evolution (via routines {\tt LUJOIN} and {\tt LUSHOW}) and hadronisation (routine {\tt LUEXEC}). In making this conversion, masses must be added to the outgoing fermions, considered massless in the hard scattering process. This is done by rescaling the fermion momenta by a single scale factor, keeping the flight directions fixed in the rest frame of the four fermion system. In the QCD evolution phase, strings join quarks coming from the same W decay. The virtuality scale of the QCD evolution is taken to be the invariant mass-squared of each evolving fermion pair. No colour reconnection is included by default, although it could be implemented by appropriate modification of routine {\tt WWGJIF} if required. Bose - Einstein correlations are neglected in the present version. The resulting event structure is then made available to the user in the {\tt /HEPEVT/} common block via a routine {\tt WWUSER} for further analysis, such as writing out for later input to a detector simulation program. The {\tt WWUSER} routine is also called at program initialisation time to allow the user to set any non-standard {\tt JETSET} program options, for example, and at termination time to allow any necessary clean-up. A dummy {\tt WWUSER} routine is supplied with the program. The only {\tt JETSET} option which is changed by {\tt WWGENPV} from its default value controls emission of gluons and photons by final-state partons\footnote{\footnotesize {\tt JETSET} parameter {\tt MSTJ(41)}}, turning off final-state photon emission simulation from {\tt JETSET} if activated in {\tt WWGENPV}, to avoid double counting. \\ All the new features of the program can be switched on/off by means of separate flags, as described in the following. \section{Input} \vskip 10pt Here we give a short explanation of the input parameters and flags required when running the program. \noindent {\bf \begin{verbatim} OGEN(CHARACTER*1) \end{verbatim}} \noindent It controls the use of the program as a Monte Carlo event generator of unweighted events ({\tt OGEN = G}) or as a Monte Carlo/adaptive integrator for weighted events ({\tt OGEN = I}). \vskip 5pt \noindent {\bf \begin{verbatim} RS(REAL*8) \end{verbatim}} \noindent The centre-of-mass energy (in GeV). \vskip 5pt \noindent {\bf \begin{verbatim} OFAST(CHARACTER*1) \end{verbatim}} \noindent It selects ({\tt OFAST = Y}) the adaptive integration branch, when {\tt OGEN = I}. When this choice is done, the required relative accuracy of the numerical integration has to be supplied by means of the {\tt REAL*8} variable {\tt EPS}. \noindent {\bf \begin{verbatim} NHITWMAX(INTEGER) \end{verbatim}} \noindent Required by the Monte Carlo integration branch. It is the maximum number of calls for the Monte Carlo loop. \vskip 5pt \noindent {\bf \begin{verbatim} NHITMAX(INTEGER) \end{verbatim}} \noindent Required by the event-generation branch. It is the maximum number of hits for the hit-or-miss procedure. \vskip 5pt \noindent {\bf \begin{verbatim} IQED(INTEGER) \end{verbatim}} \noindent This flag allows the user to switch on/off the contribution of the initial-state radiation. If {\tt IQED = 0} the distributions are computed in lowest-order approximation, while for {\tt IQED = 1} the QED corrections are included in the calculation. \vskip 5pt \noindent {\bf \begin{verbatim} OPT(CHARACTER*1) \end{verbatim}} \noindent This flag controls the inclusion of $p_T / p_L$ effects for the initial-state radiation. It is ignored in the adaptive integration branch where only initial-state strictly collinear radiation is allowed. \vskip 5pt \noindent {\bf \begin{verbatim} OFS(CHARACTER*1) \end{verbatim}} \noindent It is the option for including final-state radiation. It is assumed that final-state radiation can be switched on only if initial-state radiation including $p_T / p_L$ effects is on, in which case final-state radiation includes $p_T / p_L$ effects as well. Ignored in the adaptive integration branch. \vskip 5pt \noindent {\bf \begin{verbatim} ODIS(CHARACTER*1) \end{verbatim}} \noindent Required by the integration branch. It selects the kind of experimental distribution. For {\tt ODIS = T} the program computes the total cross section (in pb) of the process; for {\tt ODIS = W} the value of the invariant-mass distribution $d \sigma / d M$ of the system $d \bar u$ ({\tt IWCH = 1}) or of the system ${\bar d} u$ ({\tt IWCH = 2}) is returned (in pb/GeV). \vskip 5pt \noindent {\bf \begin{verbatim} OWIDTH(CHARACTER*1) \end{verbatim}} \noindent It allows a different choice of the value of the $W$-width. {\tt OWIDTH = Y} means that the tree-level Standard Model formula for the $W$-width is used; {\tt OWIDTH = N} requires that the $W$-width is supplied by the user in GeV. \vskip 15pt \vfil\eject \noindent {\bf \begin{verbatim} NSCH(INTEGER) \end{verbatim}} \noindent The value of {\tt NSCH} allows the user to choose the calculational scheme for the weak mixing angle and the gauge coupling. Three choices are available. If {\tt NSCH=1}, the input parameters used are $G_F, M_W, M_Z$ and the calculation is performed at tree level. If {\tt NSCH = 2} or {\tt 3}, the input parameters used are $\alpha(Q^2), G_F, M_W$ or $\alpha(Q^2), G_F, M_Z$, respectively, and the calculation is performed using the QED coupling constant at a proper scale $Q^2$, which is requested as further input. The recommended choice is {\tt NSCH = 2}, consistently with~\cite{wweg}. \noindent {\bf \begin{verbatim} OCOUL(CHARACTER*1) \end{verbatim}} \noindent This flag allows the user to switch on/off the contribution of the Coulomb correction. Unchanged with respect to the old version of the program. \noindent {\bf \begin{verbatim} OQCD(CHARACTER*1) \end{verbatim}} \noindent This flag allows the user to switch on/off the contribution of the naive QCD correction. \noindent {\bf \begin{verbatim} ICHANNEL(INTEGER) \end{verbatim}} \noindent A channel corresponding to a specific flavour quantum number assignment can be chosen. \noindent {\bf \begin{verbatim} ANGLMIN(REAL*8) \end{verbatim}} \noindent The minimum electron (positron) scattering angle (deg.) in the laboratory frame. It is ignored when {\tt CC20} graphs are not selected. \noindent {\bf \begin{verbatim} SRES(CHARACTER*1) \end{verbatim}} \noindent Option for switching on/off single-resonant diagrams ({\tt CC11}). \noindent {\bf \begin{verbatim} OCC20(CHARACTER*1) \end{verbatim}} \noindent Option for switching on/off single-resonant diagrams when electrons (positrons) occur in the final state ({\tt CC20}). \noindent {\bf \begin{verbatim} OOUT(CHARACTER*1) \end{verbatim}} \noindent Option for ``canonical'' output containing results for several observables and their most important moments. It is active only in the Monte Carlo integration branch. \noindent {\bf \begin{verbatim} OHAD(CHARACTER*1) \end{verbatim}} \noindent Option for switching on/off hadronisation interface in the unweighted event generation branch. \section{Test run output} \vskip 10pt The typical new calculations that can be performed with the updated version of the program are illustrated in the following examples. \vskip 8pt\noindent \leftline{\bf Sample 1} An example of adaptive integration is provided. The process considered is $e^+ e^- \to e^+ \nu_e d \bar u$ ({\tt CC20}). The output gives the cross section, together with the energy and invariant-mass losses from initial-state radiation. ``Canonical'' cuts are imposed as in~\cite{wweg}. The input card is as follows: \begin{verbatim} OGEN = I RS = 190.D0 OFAST = Y EPS = 1.D-2 IQED = 1 OPT = N OFS = N ODIS = T OWIDTH = Y NSCH = 2 ALPHM1 = 128.07D0 OCOUL = N OQCD = N ICHANNEL = 9 ANGLMIN = 10.D0 SRES = Y OCC20 = Y OOUT = N \end{verbatim} \vskip 8pt\noindent \leftline{\bf Sample 2} An example of weighted event integration is provided. Here the process considered is $e^+ e^- \to \mu^+ \nu_{\mu} d \bar u$ ({\tt CC11}). ``Canonical'' cuts are imposed as before. The ``canonical'' output is provided. The input card differs from the previous one as follows: \begin{verbatim} RS = 175.D0 OFAST = N NHITWMAX = 100000 OPT = Y OFS = Y OCOUL = Y OQCD = Y ICHANNEL = 13 OCC20 = N OOUT = Y \end{verbatim} \vskip 8pt\noindent \leftline{\bf Sample 3} An example of unweighted event generation including hadronisation is provided. A sample of 100 events corresponding to the full semileptonic channel is generated. The detailed list of an hadronised event is given. The input card differs from the first one as follows: \begin{verbatim} OGEN = G RS = 175.D0 NHITMAX = 100 OPT = Y OFS = Y OCOUL = Y OQCD = Y ICHANNEL = 26 ANGLMIN = 5.D0 OHAD = Y \end{verbatim} \vskip 10pt \vskip 15pt \section{Conclusions} The program {\tt WWGENPV 2.0} has been described. In its present version it allows the treatment of all the four-fermion reactions including the {\tt CC03} class of diagrams as a subset. This means that all the semileptonic channels are complete from the tree-level diagrams point of view, whereas fully leptonic and fully hadronic channels are treated in the CC approximation. Since the most promising channels for the $W$ mass reconstruction are the semileptonic ones, the present version of the code allows a precise analysis of such data. Moreover, NC backgrounds can be suppressed by proper invariant mass cuts, so that the code is also usable for fully hadronic and leptonic events analysis, with no substantial loss of reliability. Initial- and final-state QED radiation is taken into account within the SF formalism, including finite $p_T / p_L$ effects in the leading logarithmic approximation. The Coulomb correction is included for the {\tt CC03 } graphs. Naive QCD and leading EW corrections are implemented as well. An hadronic interface to {\tt JETSET} is also provided. The code as it stands is a valuable tool for the analysis of LEP2 data, with particular emphasis to the threshold and direct reconstruction methods for the measurement of the $W$-boson mass. Speed and high numerical accuracy allow the use of the program also for fitting purposes. The code is supported. Future releases of {\tt WWGENPV} will include: \begin{itemize} \item an interface to the code {\tt HIGGSPV}~\cite{wweg,egdp} in order to treat all the possible four-fermion processes in the massless limit, including Higgs-boson signals; \item implementation of anomalous couplings; \item implementation of CKM effects; \item the extension of the hadronic interface to {\tt HERWIG}~\cite{herwig}. \end{itemize}
proofpile-arXiv_065-438
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} The formulation of quantum field theory in terms of Haag Kastler nets of local observable algebras (``local quantum physics" \cite{Haa}) has turned out to be well suited for the investigation of general structures. Discussion of concrete models, however, is mostly done in terms of pointlike localized fields. In order to be in a precise mathematical framework, these fields might be assumed to obey the Wightman axioms \cite{StW}. Even then, the interrelation between both concepts is not yet completely understood (see \cite{BaW,BoY} for the present stage). Heuristically, Wightman fields are constructed out of Haag-Kastler nets by some scaling limit which, however, is difficult to formulate in an intrinsic way \cite{Buc2}. In a dilation invariant theory scaling is well defined, and in the presence of massless particles the construction of a pointlike field was performed in \cite{BuF}. Here, we study the possibly simplest situation: Haag-Kastler nets in 2 dimensional Min\-kows\-ki space with trivial translations in one light cone direction (``chirality") and covariant under the real M\"obius group which acts on the other lightlike direction. In \cite{FrJ}, it has been shown that in the vacuum representation pointlike localized fields can be constructed. Their smeared linear combinations are affiliated to the original net and generate it. We do not know at the moment whether they satisfy all Wightman axioms, since we have not yet found an invariant domain of definition. In \cite{Joe3}, we have generalized this to the charged sectors of a theory. We have constructed pointlike localized fields carrying arbitrary charge with finite statistics and therefore intertwining between the different superselection sectors of the theory. (In Conformal Field Theory these objects are known as ``Vertex Operators".) We have obtained the unbounded field operators as limits of elements of the reduced field bundle \cite{FRS1,FRS2} associated to the net of observables of the theory. In this paper, we start again from chiral conformal Haag-Kastler nets and present an canonical construction of N-point-functions that can be shown to fulfill the Wightman axioms. We proceed by generalizing the conformal cluster theorem \cite{FrJ} to higher N-point-functions and by examining the momentum space limit of the algebraic N-point-functions at $p=0$. We are not able to prove that these Wightman fields can be identified with the pointlike localized fields constructed in \cite{FrJ} and \cite{Joe3}. \section{First Steps} In this section, we give an explicit formulation of the setting frow which this work starts. We then present the proof of the conformal cluster theorem and the results on the construction of pointlike localized fields in \cite{FrJ} and \cite{Joe3}. \subsection{Assumptions} Let $\A=(\A(I))_{I\in\KKK}$ be a family of von Neumann algebras on some separable Hilbert space \H. \k\ denotes the set of nonempty bounded open intervals on \R. $\A$ is assumed to satisfy the following conditions. \begin{enumerate} \def\roman{enumi}){\roman{enumi})} \def\roman{enumi}{\roman{enumi}} \item Isotony: \be \A(I_1)\subset\A(I_2)\;\;\;\;\;\mb{for}\;\;\;\;I_1\subset I_2, \;\;\;\;I_1, I_2\in\k.\ee \item Locality: \be \A(I_1)\subset\A(I_2)'\;\;\;\;\;\mb{for}\;\;\;\;I_1\cap I_2=\{\}, \;\;\;\;I_1, I_2\in\k\ee ($\A(I_2)'$ is the commutant of $\A(I_2)$). \item There exists a strongly continuous unitary representation $U$ of $G=SL(2,\R)$ in $H$ with $U(-1)=1$ and \be U(g)\,\A(I)\,U(g)^{-1}=\A(gI),\;\;\;\;\;I,gI\in\k \ee ($SL(2,\R)\ni g=\left(\begin{array}{cc}a&b\\c&d\end{array}\right)$ acts on $\R\cup\{\infty\}$ by $x\mapsto \frac{ax+b}{cx+d}$ with the appropriate interpretation for $x, gx=\infty$). \item The conformal Hamiltonian \HH, which generates the restriction of $U$ to $SO(2)$, has a nonnegative spectrum. \item There is a unique (up to a phase) $U$-invariant unit vector $\OM\in\H$. \item \H\ is the smallest closed subspace containing the vacuum $\OM$ which is invariant under $U(g),$ $g\in SL(2,\R),$ and $A\in\A(I),I\in\k$ (``cyclicity"). \footnote{This assumption is seemingly weaker than cyclicity of $\OM$ w.r.t.\,the algebra of local observables on $\R$.} \end{enumerate} It is convenient to extend the net to intervals $I$ on the circle $S^1=\R\cup\{\infty\}$ by setting \be \A(I)=U(g)\;\A(g^{-1}I)\;U(g)^{-1},\;\;\;\;\;\;g^{-1}I\in\k,\;g\in SL(2,\R). \ee The covariance property guarantees that $\A(I)$ is well defined for all intervals $I$ of the form $I=gI_0,\;I_0\in\k,\;g\in SL(2,\R),$ i.e.\ for all nonempty nondense open intervals on $S^1$ (we denote the set of these intervals by \K). \subsection{Conformal Cluster Theorem} In this subsection, we derive a bound on conformal two-point-functions in algebraic quantum field theory (see \cite{FrJ}). This bound specifies the decrease properties of conformal two-point-functions in the algebraic framework to be exactly those known from theories with pointlike localization. The Conformal Cluster Theorem plays a central role in this work. \medskip {\bf Conformal Cluster Theorem (see \cite{FrJ}):} Let $(\A(I))_{I\in\KKK}$ be a conformally covariant local net on $\R$. Let $a,b,c,d\in\R$ and $a<b<c<d$. Let $A\in\A(\,(a,b)\,)$, $B\in\A(\,(c,d)\,)$, $n\in\N$ and $P_k A\OM=P_k A^*\OM=0,\;k<n$. $P_k$ here denotes the projection on the subrepresentation of $U(G)$ with conformal dimension $k$. We then have \be |(\,\OM,BA\OM\,)|\leq \left(\frac{(b-a)\,(d-c)}{(c-a)\,(d-b)}\right)^n \;\|A\|\,\|B\|. \ee \medskip {\bf Proof:} Choose $R>0$. We consider the following 1-parameter subgroup of $G=SL(2,\R)$\,: \be g_t\,:\,x\longmapsto\frac{x\, \mb{cos}\frac{t}{2}+R\,\mb{sin}\frac{t}{2}} {-\frac{x}{R}\,\mb{sin}\frac{t}{2}+\mb{cos}\frac{t}{2}}\,. \ee Its generator ${\rm \bf H}_R$ is within each subrepresentation of $U(G)$ unitarily equivalent to the conformal Hamiltonian ${\rm \bf H}$. Therefore, the spectrum of $A\OM$ and $A^*\OM$ w.r.t.\ ${\rm \bf H}_R$ is bounded below by $n$. Let $0<t_0<t_1<2\pi$ such that $g_{t_0}(b)=c$ and $g_{t_1}(a)=d$. We now define \be F(z)=\left\{\begin{array}{ll} (\,\OM,\,B\,z^{-{\rm \bf H}_R}\,A\OM\,) &|z|>1\\ (\,\OM,\,A\,z^{{\rm \bf H}_R}\,B\OM\,)&|z|<1\\ (\,\OM,\,A\,\alpha_{g_t}(B)\,\OM\,)&z=e^{it},\,t\not\in [t_0,t_1] \end{array}, \right. \ee a function analytic in its domain of definition, and then \be G(z)=\,(z-z_0)^n\,(z^{-1}-z_0^{-1})^n\,F(z),\;\; z_0=e^{\frac{i}{2}(t_0+t_1)}\,. \ee (Confer the idea in \cite{Fre}.) At $z=0$ and $z=\infty$ the function $G(\cdot)$ is bounded because of the bound on the spectrum of ${\rm \bf H}_R$ and can therefore be analytically continued. As an analytic function it reaches its maximum at the boundary of its domain of definition, which is the interval $[e^{it_0},e^{it_1}]$ on the unit circle: \be \mb{sup}|G(z)|\,\leq\,\|A\|\,\|B\|\,|e^{it_0}-e^{\frac{i}{2} (t_0+t_1)}|^{2n}\,=\,\|A\|\,\|B\|\,(2\, \mb{sin}\frac{t_0-t_1}{4})^{2n} \,. \ee This leads to \beam |(\,\OM,\,BA\OM\,)|&=&|F(1)|\,=\,|G(1)|\ |1-e^{\frac{i}{2}(t_0+t_1)} |^{ -2n}\,=\,|G(1)|\ (2\,\mb{sin}\frac{t_0+t_1}{4})^{-2n}\nn\\ &\leq&\mb{sup}|G|\,(2\,\mb{sin}\frac{t_0+t_1}{4})^{-2n}\,\leq\, \|A\|\,\|B\|\,\left(\frac{\mb{sin}\frac{t_0-t_1}{4}}{\mb{sin} \frac{t_0+t_1}{4}}\right)^{2n}\,. \eeam Determining $t_0$ and $t_1$ we obtain \be \lim_{R\rightarrow\infty}R\,t_0=2(c-b)\;\;\;\mb{and}\;\;\; \lim_{R\rightarrow\infty}R\,t_1=2(d-a)\,. \ee We now assume $a-b=c-d$ and find $\left(\frac{t_0-t_1}{t_0+t_1} \right)^2=\frac{(a-b)\,(c-d)}{(a-c)\,(b-d)}=:x\,.$ Since the bound on $|(\,\OM,\,BA\OM\,)|$ can only depend on the conformal cross ratio $x$, we can drop the assumption and the theorem is proven.\,\hfill $\Box$ \subsection{The Construction of Pointlike Localized Fields from Conformal Haag-Kastler Nets} This subsection presents the argumentation and results of \cite{FrJ} and \cite{Joe3}:\\ The idea for the definition of conformal fields is the following: Let $A$ be a local observable, \be A\in\bigcup_{I\in\KKK}\A(I), \ee and $P_\tau$ the projection onto an irreducible subrepresentation $\tau$ of $U$. The vector $P_\tau A\OM$ may then be thought of as $\varphi_\tau(h)\,\OM$ where $\varphi_\tau$ is a conformal field of dimension $n_\tau=:n$ and $h$ is an appropriate function on $\R$. The relation between $A$ and $h$, however, is unknown at the moment, up to the known transformation properties under $G$, \be U(g)\,P_\tau A\OM=\varphi_\tau(h_g^{(n)})\,\OM \ee with $h_g^{(n)}(x)=(cx-a)^{2n-2}\,h(\frac{dx-b}{-cx+a})$, $g=\left(\begin{array}{cc}a&b\\c&d\end{array}\right)\in G$. We may now scale the vector $P_\tau A\OM$ by dilations $D(\l)= U\left(\begin{array}{cc}\l^{\frac{1}{2}}&0\\0&\l^{-\frac{1}{2}} \end{array}\right)$ and find \be D(\l)\,P_\tau A\OM=\l^n\,\varphi_\tau(h_{\l})\,\OM \ee where $h_{\l}(x)=\l^{-1}\,h(\frac{x}{\lambda })$. Hence, we obtain formally for $\lambda\downarrow 0$ \be \l^{-n}\,D(\l)\,P_\tau A\OM\longrightarrow \int dx\,h(x)\;\varphi_\tau(0) \,\OM. \ee In order to obtain a Hilbert space vector in the limit, we smear over the group of translations $T(a)=U\left(\begin{array}{cc}1&a\\0&1\end{array} \right)$ with some test function $f$ and obtain formally \be \label{a} \lim_{\lambda \downarrow 0}\l^{-n}\int da\,f(a)\;T(a)\,D(\l)\,P_\tau A\OM= \int dx\,h(x)\;\varphi_\tau(f)\,\OM.\label{f} \ee We now interpret the left-hand side as a definition of a conformal field $\varphi_\tau$ on the vacuum, and try to obtain densely defined operators with the correct localization by defining \be \varphi_\tau^I(f)\,A'\OM=A'\varphi_\tau^I(f)\,\OM,\;\; f\in\D(I),\,A'\in\A(I)',\,I\in\K. \ee In order to make this formal construction meaningful, there are two problems to overcome. The first one is the fact that the limit on the left-hand side of (\ref{a}) does not exist in general if $A\OM$ is replaced by an arbitrary vector in $H$. This corresponds to the possibility that the function $h$ on the right-hand side might not be integrable. We will show that after smearing the operator $A$ with a smooth function on $G$, the limit is well defined. Such operators will be called regularized. The second problem is to show that the smeared field operators $\varphi_\tau^I(f)$ are closable, in spite of the nonlocal nature of the projections $P_\tau$. We omit the technical parts of \cite{FrJ} and \cite{Joe3} and summarize the results in a compact form and as general as possible. Due to the positivity condition, the representation $U(\tilde{G})$ is completely reducible into irreducible subrepresentations and the irreducible components $\tau$ are up to equivalence uniquely characterized by the conformal dimension $n_\tau\in\Rp$ ($n_\tau$ is the lower bound of the spectrum of the conformal Hamiltonian \HH\ in the representation $\tau$). Associated with each irreducible subrepresentation $\tau$ of $U$ we find for each $I\in\k$ a densely defined operator-valued distribution $\varphi_\tau^I$ on the space $\D(I)$ of Schwartz functions with support in $I$ such that the following statements hold for all $f\in\D(I).$ \begin{enumerate} \def\roman{enumi}){\roman{enumi})} \def\roman{enumi}{\roman{enumi}} \item The domain of definition of $\varphi_\tau^I(f)$ is given by $\A(I')\,\OM$. \item \be \varphi_\tau^I(f)\,\OM\in P_\tau\H_{red} \ee with $P_\tau$ denoting the projector on the module of $\tau$. \item \be U(\tilde{g})\; \varphi_\tau^I(x)\; U(\tilde{g})^{-1}=(cx+d)^{-2n_\tau}\varphi_\tau^{gI} (\tilde{g}x) \ee with the covering projection $\tilde{g}\mapsto g$ and $g=\left(\begin{array}{cc}a&b\\c&d\end{array}\right)\in SL(2,\R),\;I, gI\in\k$. \item $\varphi_\tau^I(f)$ is closable. \item The closure of $\varphi_{\tau}^I(f),\,f\in\D(I),$ is affiliated to $\A(I)$. \item $\A(I)$ is the smallest von Neumann algebra to which all operators $\varphi_{\tau}^I(f),\,f\in\D(I),$ are affiliated. \item The exchange algebra of the reduced field bundle \cite{FRS2} and the existence of the closed field operators $\varphi_\tau^I(f)$, mapping a dense set of the vacuum Hilbert space into some charged sector with finite statistics, suffice to construct closed field operators $\varphi_{\tau,\alpha}^I(f)$, mapping a dense set of an arbitrary charged sector $\alpha$ with finite statistics into some (other) charged sector with finite statistics. Here, the irreducible module $\tau$ of $U(\tilde{G})$ labels orthogonal irreducible fields defined in the same sector $\alpha$\,. \item The closure of any $\varphi_{\tau,\alpha}^I(f),\,f\in\D(I),$ is affiliated to $\F_{red}(I)$. \item $\F_{red}(I)$ is the smallest von Neumann algebra to which all operators $\varphi_{\tau,\alpha}^I(f),\,f\in\D(I),$ are affiliated. \end{enumerate} With the existence of pointlike localized fields we are able to give a proof of a generalized Bisognano-Wichmann property. We can identify the conformal group and the reflections as generalized modular structures in the reduced field bundle. Especially, we obtain a PCT operator on $\H_{red}$ proving the PCT theorem for the full theory. Moreover, the existence of pointlike localized fields gives a proof of the hitherto unproven Spin-Statistics theorem for conformal Haag-Kastler nets in 1+1 dimensions. It was also possible to prove an operator product expansions for arbitrary local observables: \\ For each $I\in\k$ and each $A\in \A(I)$ there is a local expansion \be A\,=\,\sum_{\tau}\varphi_{\tau}^I(f_{\tau,A})\, \ee into a sum over all irreducible modules $\tau$ of $U(G)$ with \be \mb{supp}f_{\tau,A}\subset I\,,\ee which converges on $\A(I')\OM$ $*$-strongly (cf.\ the definition in \cite{BrR}). Here, $I'$ denotes the complement of $I$ in $\k$. \section{Canonical Construction of Wightman Fields} Starting from a chiral conformal Haag-Kastler net, pointlike localized fields have been constructed in \cite{FrJ,Joe3}. Their smeared linear combinations are affiliated to the original net and generate it. We do not know at the moment whether these fields satisfy all Wightman axioms, since we have not found an invariant domain of definition. In this section, we construct in a canonical manner a complete set of pointlike localized correlation functions out of the net of algebras we have been starting from. We proceed by generalizing the conformal cluster theorem to higher N-point-functions and by examining the momentum space limit of the algebraic N-point-functions at $p=0$. This canonically constructed set of correlation functions can be shown to fulfill the conditions for Wightman functions (cf.\ \cite{StW} and \cite{Jos}). Hence, we can construct an associated field theory fulfilling the Wightman axioms. We are not able to prove that these Wightman fields can be identified with the pointlike localized fields constructed in \cite{FrJ} and \cite{Joe3}. We do not know either how the Haag-Kastler theory, we have been starting from, can be reconstructed from the Wightman theory. Such phenomena have been investigated by Borchers and Yngvason \cite{BoY}. Starting from a Wightman theory, they could not rule out in general the possibility that the associated local net has to be defined in an enlarged Hilbert space. \subsection{Conformal Two-Point-Functions} First, we will determine the general form of conformal two-point-functions of local observables:\\ It has been shown (cf.\ e.g.\ \cite{Joe1}) that a two-point-function $(\,\OM,\,B\,U(x)\,A\OM\,)$ of a chiral local net with translation covariance is of Lebesgue class $L^p$ for any $p>1$. The Fourier transform of this two-point-function is a measure concentrated on the positive half line. Therefore, it is - with the possible exception of a trivial delta function at zero - fully determined by the Fourier transform of the commutator function $(\,\OM,\, [B,\,U(x)\,A\,U(x)^{-1}]\,\OM\,)\,.$ Since $A$ and $B$ are local observables, the commutator function has compact support and an analytic Fourier transform $G(p)$. The restriction $\Theta(p)\,G(p)$ of this analytic function to the positive half line is then the Fourier transform of $(\,\OM,\,B\,U(x)\,A\OM\,)\,.$ In the conformally covariant case with $P_kA\OM=P_kA^*\OM=0,\,k<n$, the conformal cluster theorem implies that the two-point-function $(\,\OM,\,B\,U(x)\,A\OM\,)$ decreases as $x^{-2n}$. Therefore, its Fourier transform is $2n\!-\!2\,$times continuously differentiable and can be written as $\Theta(p)\,p^{2n-1}\,H(p)$ with an appropriate analytic function $H(p)$. Using this result, we are able to present a sequence of canonically scaled two-point-functions of local observables converging as distributions to the two-point-function known from conventional conformal field theory (cf.\ \cite{Joe1,Reh}): \be \lim_{\lambda\downarrow 0}\l^{-2n}\ (\,\OM,\,B\,U( \lambda^{-1}x)\,A\OM\,)\ =\ \lim_{\lambda\downarrow 0}\l^{-2n}\ {\cal F}_{p\rightarrow x}\ \Theta(p)\,(\l p)^{2n-1}\,H(\l p)\,\l\,dp\ =\ H(0)\ (x+i\varepsilon)^{-2n}\,. \ee \subsection{Conformal Three-Point Functions} We consider the properties of chiral algebraic three-point functions \be (\,\OM,\,A_1\,U(x_1-x_2)\,A_2\,U(x_2-x_3)\,A_3\,\OM\,) \ee of local observables $A_i$\,, $ i=1,2,3$\,. The general form of a (truncated) chiral three-point function of local observables is restricted by locality and by the condition of positive energy. The Fourier transform of an algebraic three-point function can be shown to be the sum of the restrictions of analytic functions to disjoint open wedges in the domain of positive energy: \\ If $F$ now denotes the Fourier transform of $(\,\OM,\,A_1\,U(\cdot)\,A_2\,U(\cdot)\,A_3\,\OM\,)$, we get by straightforward calculations as a first result (cf.\ \cite{Joe4}) \be F(p,q)\,=\,\Theta(p)\,\Theta(q-p)\,G^+(p,q)\,+\, \Theta(q)\,\Theta(p-q)\,G^-(p,q) \ee with appropriate analytic functions $G^+$ and $G^-$. In the case of conformal covariance the general form of these algebraic three-point functions is even more restricted by the following generalization of the conformal cluster theorem \cite{FrJ}: \medskip {\bf Theorem:} Let $(\A(I))_{I\in\KKK}$ be a conformally covariant local net on $\R$\,. Let $a_i,b_i\in\R\,,\ i=1,2,3$\,, and $a_1<b_1<a_2<b_2<a_3<b_3$\,. Let $A_i\in\A(\,(a_i,b_i)\,)$\,, $n_i\in\N$\,, $i=1,2,3$\,, and \be P_k\,A_i\,\OM=P_k\,A_i^*\,\OM=0\,,\;k<n_i\,. \ee $P_k$ here denotes the projection on the subrepresentation of $U(SL(2,R))$ with conformal dimension $k$\,. We then have the following bound: \beam |(\,\OM,A_1A_2A_3\OM\,)|&\leq& \left|\frac{(a_1-b_1)+(a_2-b_2)}{(a_2-a_1)+(b_2-b_1)}\right|^{(n_1+n_2-n_3)}\\ && \left|\frac{(a_1-b_1)+(a_3-b_3)}{(a_3-a_1)+(b_3-b_1)}\right|^{(n_1+n_3-n_2)} \;\;\nn\\&& \left|\frac{(a_2-b_2)+(a_3-b_3)}{(a_3-a_2)+(b_3-b_2)}\right|^{(n_2+n_3-n_1)} \;\|A_1\|\,\|A_2\|\,\|A_3\|\,.\nn \eeam If we additionally assume \be a_1-b_1=a_2-b_2=a_3-b_3\,, \ee we get \be |(\,\OM,A_1A_2A_3\OM\,)|\ \leq\ r_{12}^{(n_1+n_2-n_3)/2}\,r_{23}^{(n_2+n_3-n_1)/2}\, r_{13}^{(n_1+n_3-n_2)/2}\ \|A_1\|\,\|A_2\|\,\|A_3\|\,, \ee with the conformal cross ratios \be \frac{(a_i-b_i)\,(a_j-b_j)}{(a_i-a_j)\,(b_i-b_j)}=:r_{ij}\,,\,i,j=1,2,3\,. \ee \medskip {\bf Proof:} This proof follows, wherever possible, the line of argument in the proof of the conformal cluster theorem for two-point functions (cf.\ \cite{FrJ}). \\ Choose $R>0$\,. Let us consider the following one-parameter subgroup of $SL(2,\R)$\,: \be g_t\,:\,x\longmapsto\frac{x\, \mb{cos}\frac{t}{2}+R\,\mb{sin}\frac{t}{2}} {-\frac{x}{R}\,\mb{sin}\frac{t}{2}+\mb{cos}\frac{t}{2}}\,. \ee Its generator ${\rm \bf H}_R$ is within each subrepresentation of $U(SL(2,R))$ unitarily equivalent to the conformal Hamiltonian ${\rm \bf H}$\,. Therefore, the spectrum of $A_i\,\OM$ and $A^*_i\,\OM$ with respect to ${\rm \bf H}_R$ is bounded from below by $n_i$\,, $i=1,2,3$\,. Let $0<t^-_{ij}<t^+_{ij}<2\pi$ such that \be g_{t^-_{ij}}(b_i)=a_j \ee and \be g_{t^+_{ij}}(a_i)=b_j\, \ee for $i,j=1,2,3$\,, $i<j$\,. We now define \be F(z_1,z_2,z_3):=(\,\OM,\,A_{i_1}\,(\frac{z_{i_1}}{z_{i_2}})^{{\rm \bf H}_R}\,A_{i_2}\,(\frac{z_{i_2}}{z_{i_3}})^{{\rm \bf H}_R}\,A_{i_3}\OM\,) \ee in a domain of definition given by \be |z_{i_1}|<|z_{i_2}|<|z_{i_3}| \ee with permutations $(i_1,i_2,i_3)$ of $(1,2,3)$\,. This definition can uniquely be extended to certain boundary values with $|z_{j}|=|z_{k}|$\,, $j,k=1,2,3$\,, $j\neq k$\,:\\ $F$ shall be continued to this boundary of its domain of definition if \be t_{jk}:=-i\log\frac{z_{j}}{z_{k}}\notin[t^-_{jk},t^+_{jk}]+2\pi{\bf Z} \ee or equivalently if \be g_{t_k}([a_k,b_k])\cap g_{t_j}([a_j,b_j])\neq\emptyset\,, \ee using the notation \be t_i:=-i\log z_i\,,\ i=1,2,3\,. \ee Thereby, boundary points with coinciding absolute values are included in the domain of definition. The definition of $F$ is chosen in analogy to the analytic continuation of general Wightman functions (cf., e.g., \cite{StW,Jos}) such that the edge-of-the-wedge theorem for distributions with several variables \cite{StW} proves $F$ to be an analytic function:\\ Permuting the local observables $A_i$\,, $i=1,2,3$\,, we have six three-point functions \be (\,\OM,\,A_{i_1}\,U(x_{i_1}-x_{i_2})\,A_{i_2}\,U(x_{i_2}-x_{i_3})\,A_{i_3}\OM\,)\,. \ee These six functions have by locality identical values on a domain \be E:=\{(y_1,y_2)\in {\bf R}^2\,|\,|y_1|>c_1,\,|y_2|>c_2,\,|y_1+y_2|>c_3\} \ee with appropriate $c_1,c_2,c_3\in\Rp$\,. Each single function can be continued analytically by the condition of positive energy to one of the six disjoint subsets in \be U:={\bf R}^2+iV:=\{(z_1,z_2)\in {\bf C}^2\, |\,\mb{Im}z_1\neq 0 \neq \mb{Im}z_2\,,\,\mb{Im}z_1 +\mb{Im}z_2\neq 0\}\,. \ee In this geometrical situation, the edge-of-the-wedge theorem (cf.\ \cite{StW}, theorem 2.14) proves the assumed analyticity of $F$.\\ With the abbreviation \be z_{ij}^0:=e^{i(t^-_{ij}+t^+_{ij})/2}\,,\ i,j=1,2,3\,, \ee we then define \beam \lefteqn{ G(z_1,z_2,z_3)}\\ &:=&F(z_1,z_2,z_3)\,\prod_{(i,j,k)\in T(1,2,3)}\,(\frac{z_i}{z_j}-z^0_{ij})^{(n_i+n_j-n_k)/2}\,(\frac{z_j}{z_i}-z^0_{ji})^{(n_i+n_j-n_k)/2}\,,\nn \eeam where $T(1,2,3)$ denotes the set $\{(1,2,3),\,(1,3,2),\,(2,3,1)\}$\,. The added polynomial in $z_i$\,, $i=1,2,3$\,, is constructed such that the degree of the leading terms are restricted by the assumption on the conformal dimensions of the three-point function $F$. Also, using the binomial formula, it can be controlled by straightforward calculations that no half odd integer exponents appear after multiplication of the product. Hence, at $z_i=0$ and $z_i=\infty$\,, $i=1,2,3$\,, the function $G$ is bounded because of the bound on the spectrum of ${\rm \bf H}_R$ and can therefore be analytically continued. We can find estimates on $G$ by the maximum principle for analytic functions. In order to get the estimate needed in this proof, we do not use the maximum principle for several complex variables \cite{BoM}. Instead, we present an iteration of the maximum principle argument used in the proof of the conformal cluster theorem \cite{FrJ} for the single variables $z_i$\,, $i=1,2,3$\,, of $G(\cdot,\cdot,\cdot)$ and derive a bound on $G(1,1,1)$\,:\\ Applying the line of argument known from the case of the two-point functions now to $G(\cdot,1,1)$\,, we get the estimate \beam |G(1,1,1)|&\leq& \mbox{sup}_{z_1}\,|G(z_1,1,1)|\nn\\ &=& \mbox{sup}_{z_1\in B_{\cdot,1,1}}\,|G(z_1,1,1)|\,. \eeam The boundary of the domain of definition of the maximal analytical continuation of $G(\cdot,1,1)$ is here denoted by \be B_{\cdot,1,1}:=\{e^{it}\,|\,t\notin [t_{12}^-,t_{12}^+]\cup [t_{13}^-,t_{13}^+]+2\pi{\bf Z}\}\,. \ee Applying this argument to $G(z_1,\cdot,1)$, we analogously get the estimate \beam |G(z_1,1,1)|&\leq& \mbox{sup}_{z_2}\,|G(z_1,z_2,1)|\nn\\ &=& \mbox{sup}_{z_2\in B_{z_1,\cdot,1}}\,|G(z_1,z_2,1)|\, \eeam with $B_{z_1,\cdot,1}$ denoting the boundary of the domain of definition of the maximal analytical continuation of $G(z_1,\cdot,1)$\,. Applying this argument finally to $G(z_1,z_2,\cdot)$\,, we analogously get the estimate \beam |G(z_1,z_2,1)|&\leq& \mbox{sup}_{z_3}\,|G(z_1,z_2,z_3)|\nn\\ &=& \mbox{sup}_{z_3\in B_{z_1,z_2,\cdot}}\,|G(z_1,z_2,z_3)|\, \eeam with $B_{z_1,z_2,\cdot}$ denoting the boundary of the domain of definition of the maximal analytical continuation of $G(z_1,z_2,\cdot)$\,. Having iterated this maximum principle argument for the single variables $z_i$\,, $i=1,2,3$\,, we can combine the derived estimates and get \be |G(1,1,1)|\ \leq\ \mbox{sup}_{t_{jk}=-i\log\frac{z_j}{z_k}\notin[t^-_{jk},t^+_{jk}]+2\pi{\bf Z}\,,\ j\neq k}\,|G(z_1,z_2,z_3)|\,. \ee Hence, the boundary values of $G$ have to be evaluated on the domain described by \be g_{t_k}([a_k,b_k])\cap g_{t_j}([a_j,b_j])\neq\emptyset\, \ee with $ t_i=-i\log z_i\,,\ i=1,2,3\,. $ We find the supremum with the same calculation as in the proof of the conformal cluster theorem above (cf.\ \cite{FrJ}): \beam |G(1,1,1)|\,&\leq&\,\|A_1\|\,\|A_2\|\,\|A_3\|\,\prod_{(i,j,k)\in T(1,2,3)}\,|e^{it^-_{ij}}-e^{i(t^-_{ij}+t^+_{ij})/2}|^{n_i+n_j-n_k}\nn\\&=&\,\|A_1\|\,\|A_2\|\,\|A_3\|\,\prod_{(i,j,k)\in T(1,2,3)}\,|2\, \mb{sin}\frac{t^-_{ij}-t^+_{ij}}{4}|^{n_i+n_j-n_k} \eeam This leads to another estimate: \beam |(\,\OM,A_1A_2A_3\OM\,)|&=&\,|F(1,1,1)|\nn\\ &=&\,|G(1,1,1)|\,\prod_{(i,j,k)\in T(1,2,3)}\,|1-e^{i(t^-_{ij}+t^+_{ij})/2}|^{n_i+n_j-n_k}\nn\\ &=&\,|G(1,1,1)|\,\prod_{(i,j,k)\in T(1,2,3)}\,|2\,\mb{sin}\frac{t^-_{ij}+t^+_{ij}}{4}|^{n_i+n_j-n_k}\nn\\ &\leq&\,\|A_1\|\,\|A_2\|\,\|A_3\|\,\prod_{(i,j,k)\in T(1,2,3)}\,\left|\frac{\mb{sin}\frac{t^-_{ij}-t^+_{ij}}{4}}{\mb{sin} \frac{t^-_{ij}+t^+_{ij}}{4}}\right|^{n_i+n_j-n_k} \eeam Determining $t^-_{ij}$ and $t^+_{ij}$\,, we obtain for $i,j=1,2,3$ \be \lim_{R\rightarrow\infty}R\,t^-_{ij}=2(a_j-b_i) \ee and \be \lim_{R\rightarrow\infty}R\,t^+_{ij}=2(b_j-a_i) \ee and the first bound in the theorem is proven. If we now assume \be a_1-b_1=a_2-b_2=a_3-b_3\,, \ee we find \be \left(\frac{t^-_{ij}-t^+_{ij}}{t^-_{ij}+t^+_{ij}} \right)^2=\frac{(a_i-b_i)\,(a_j-b_j)}{(a_i-a_j)\,(b_i-b_j)}=r_{ij}\,,\;\;\;\,i,j=1,2,3\,, \ee and the theorem is proven.\hfill $\Box$ \medskip This theorem can be used to get deeper insight in the form of the Fourier transforms of algebraic three-point functions. As in the case of the two-point functions, we proceed by transferring the decrease properties of the function in position space into regularity properties of the Fourier transform in momentum space. In conventional conformal field theory, the three-point function with conformal dimensions $n_i\,,\ i=1,2,3$\,, is known up to multiplicities as \beam f_{n_1n_2n_3}(x_1,x_2,x_3) &=&(x_1-x_2+i\varepsilon)^{-(n_1+n_2-n_3)}\nn\\ &&(x_2-x_3+i\varepsilon)^{-(n_2+n_3-n_1)}\nn\\ &&(x_1-x_3+i\varepsilon)^{-(n_1+n_3-n_2)} \eeam (cf.\ \cite{ChH,Reh}). Its Fourier transform \be \tilde{f}_{n_1n_2n_3}(p,q)\,=:\,\Theta(p)\,\Theta(q)\,Q_{n_1n_2n_3}(p,q) \ee can be calculated to be a sum of the restrictions of homogeneous polynomials $Q^+_{n_1n_2n_3}$ and $Q^-_{n_1n_2n_3}$ of degree $n_1+n_2+n_3-2$ to disjoint open wedges $W_+$ and $W_-$ in the domain of positive energy (cf.\ \cite{Reh}). By the bound in the cluster theorem above, we know that a conformally covariant algebraic three-point function $(\,\OM,\,A_1\,U(x_1-x_2)\,A_2\,U(x_2-x_3)\,A_3\OM\,)$ of local observables $A_i$ with minimal conformal dimensions $n_i$\,, $i=1,2,3\,,$ decreases in position space at least as fast as the associated pointlike three-point function $f_{n_1n_2n_3}(x_1,x_2,x_3)$ known from conventional conformal field theory. Hence, the Fourier transform $F_{A_1A_2A_3}(p,q)$ of this algebraic three-point function has to be at least as regular in momentum space as the Fourier transform $\tilde{f}_{n_1n_2n_3}(p,q)$ of the associated pointlike three-point function known from conventional conformal field theory: \\ Technically, we use a well-known formula from the theory of Fourier transforms, \be {\cal F}(\mbox{Pol}(X)S)=\mbox{Pol}(\frac{\partial}{\partial Y}){\cal F}S\,, \ee for arbitrary temperate distributions $S$ and polynomials Pol$(\cdot)$ with a (multi-dimensional) variable $X$ in position space and an appropriate associated differential operator $\frac{\partial}{\partial Y}$ in momentum space. $\F$ denotes the Fourier transformation from position space to momentum space. Let now $S$ be the conformally covariant algebraic three-point function of local observables $A_i$ with minimal conformal dimensions $n_i$\,, $i=1,2,3\,$: \be S\,:=\,(\,\OM,\,A_1\,U(x_1-x_2)\,A_2\,U(x_2-x_3)\,A_3\,\OM\,)\, \ee and $X$ be a pair of two difference variables out of $x_i-x_j$\,, $i,j=1,2,3\,.$ By the cluster theorem proved above, we can now choose an appropriate homogeneous polynomial $\mbox{Pol}(X)$ of degree $n_1+n_2+n_3-4$ such that the product $\mbox{Pol}(X)\,S$ is still absolutely integrable in position space. Using the formula given above, we see that $\mbox{Pol}(\frac{\partial}{\partial Y}){\cal F}S$ is continuous and bounded in momentum space. Furthermore, we have already derived the form of the Fourier transform $F$ of an arbitrary (truncated) algebraic three-point function in a chiral theory to be \be F(p,q)\,=\,\Theta(p)\,\Theta(q-p)\,G^+(p,q)\,+\, \Theta(q)\,\Theta(p-q)\,G^-(p,q) \ee with appropriate analytic functions $G^+$ and $G^-$. Thereby, we see that in the case of conformal covariance with minimal conformal dimensions $n_i$\,, $i=1,2,3$\,, the analytic function $G^+$ ($G^-$) can be expressed as the product of an appropriate homogeneous polynomial $P^+$ ($P^-$) of degree $n_1+n_2+n_3-2$ restricted to the wedge $W_+$ ($W_-$) and an appropriate analytic function $H^+$ ($H^-$)\,. Hence, we have proved that the Fourier transform $F_{A_1A_2A_3}$ of the algebraic three-point function $(\,\OM,\,A_1\,U(x_1-x_2)\,A_2\,U(x_2-x_3)\,A_3\,\OM\,)$ can be written as \be F_{A_1A_2A_3}(p,q)\ =\ \Theta(p)\,\Theta(q)\,P_{A_1A_2A_3}(p,q)\,H_{A_1A_2A_3}(p,q) \ee with an appropriate homogeneous function $P_{A_1A_2A_3}(p,q)$ of degree $n_1+n_2+n_3-2$ and an appropriate continuous and bounded function $H_{A_1A_2A_3}(p,q)$\,. These results suffice to control the pointlike limit of the considered correlation functions. Scaling an algebraic three-point function in a canonical manner, we construct a sequence of distributions that converges to the three-point function of conventional conformal field theory: \beam &&\nn\\ \lefteqn{ \lim_{\lambda\downarrow 0}\l^{-(n_1+n_2+n_3)}\ (\,\OM,\,A_1\,U( \frac{x_1-x_2}{\lambda})\,A_2\,U( \frac{x_2-x_3}{\lambda})\,A_3\,\OM\,)}\nn\\ &&\nn\\ &=&\lim_{\lambda\downarrow 0}\l^{-(n_1+n_2+n_3)} \ {\cal F}_{p\rightarrow x_1-x_2\atop q\rightarrow x_2-x_3}\ F_{A_1A_2A_3}(\l p,\l q)\,\l^2\,dp\,dq\nn\\ &&\nn\\ &=&\lim_{\lambda\downarrow 0}\l^{-(n_1+n_2+n_3)}\nn\\ && \ {\cal F}_{p\rightarrow x_1-x_2\atop q\rightarrow x_2-x_3}\ \Theta(p)\,\Theta(q)\, \l^{n_1+n_2+n_3-2}\,P_{A_1A_2A_3}(p,q)\,H_{A_1A_2A_3}(\l p,\l q)\,\l^2\,dp\,dq\nn\\ &&\nn\\ &=&(x_1-x_2+i\varepsilon)^{-(n_1+n_2-n_3)}\nn\\ &&(x_2-x_3+i\varepsilon)^{-(n_2+n_3-n_1)}\nn\\ &&(x_1-x_3+i\varepsilon)^{-(n_1+n_3-n_2)}\ H_{A_1A_2A_3}(0,0)\,.\\ &&\nn \eeam \subsection{Conformal N-Point Functions} Since the notational expenditure increases strongly as we come to the construction of higher N-point functions, we concentrate on qualitatively new aspects not occurring in the case of two-point functions and three-point functions. These qualitatively new aspects in the construction of higher N-point functions are related to the fact that in conventional field theory the form of higher N-point functions is not fully determined by conformal covariance. In conventional conformal field theory conformal covariance restricts the form of correlation functions of field operators $\varphi_i(x_i)\,,\ i=1,2,...\,,N\,,$ with conformal dimension $n_i$ in the following manner (cf.\ \cite{ChH,Reh}): \be (\OM,\,\left(\prod_{1\leq i\leq N}\,\varphi_i(x_i)\right)\,\OM)\ =\ \left(\prod_{1\leq i<j\leq N}\frac{1}{(x_j-x_i+i\e)^{c_{ij}}}\right)\;f(r_{t_1u_1}^{v_1s_1},...\,,r_{t_{N-3}u_{N-3}}^{v_{N-3}s_{N-3}})\,.\ee Here, $f(\cdot,...\,,\cdot)$ denotes an appropriate function depending on $N\!-\!3$ algebraicly independent conformal cross ratios \be r_{tu}^{vs}:=\frac{(x_v-x_s)}{(x_v-x_t)}\frac{(x_t-x_u)}{(x_s-x_u)}\,. \ee The exponents $c_{ij}$ must fulfill the consistency conditions \be \sum_{j=1\atop j\neq i}^N c_{ij}=2n_i\,,\ c_{ij}=c_{ji}\,,\ 1\leq i\leq N\,. \ee These conditions do not fully determine the exponents $c_{ij}$ in the case of $N\!\geq\!4$\,. Hence, in conventional conformal field theory four-point functions and higher N-point functions are not fully determined by conformal covariance. In the case of conformal two-point functions and conformal three-point functions, our strategy to construct pointlike localized correlation functions was the following: First, we proved that the algebraic correlation functions decrease in position space as fast as the associated correlation functions in conventional field theory, which are uniquely determined by conformal covariance. Then, we transferred this property by Fourier transformation into regularity properties in momentum space. Finally, we were able to prove that the limit $\lambda\!\downarrow\!0$ of canonically scaled algebraic correlation functions converges to (a multiple of) the associated pointlike localized correlation functions in conventional conformal field theory. In the case of four-point functions and higher N-point functions, the situation has changed and we cannot expect to be able to fully determine the form of the pointlike localized limit in this construction, since for $N>4$ the correlation functions in conventional conventional field theory are not any longer uniquely determined by conformal covariance. Beginning with the discussion of the general case with $N\!\geq\!4$\,, we consider algebraic N-point functions \be (\,\OM,\,\left(\prod_{1\leq i\leq N}\,U(-x_i)\,A_i\,U(x_i)\right)\,\OM\,) \ee of local observables $A_i$ with minimal conformal dimensions $n_i$\,, $i=1,2,...\,,N$\,, in a chiral theory with conformal covariance. We want to examine the pointlike limit of canonically scaled correlation functions \be \lim_{\lambda\downarrow 0}\l^{-\left(\sum_{1\leq i\leq N}n_i\right)}\ (\,\OM,\,\left(\prod_{1\leq i\leq N}\,U(- \frac{x_i}{\lambda})\,A_i\,U( \frac{x_i}{\lambda})\right)\,\OM\,)\,. \label{i} \ee Our procedure in the construction of pointlike localized N-point functions for $N\!\geq\!4$ will be the following: We consider all possibilities to form a set of exponents $c_{ij}$ fulfilling the consistency conditions \be \sum_{j=1\atop j\neq i}^N c_{ij}=2n_i\,,\ c_{ij}=c_{ji}\,,\ i=1,2,3,...\,,N\,. \ee For each consistent set of exponents a bound on algebraic N-point functions in position space can be proved. Each single bound on algebraic N-point functions in position space can be transferred into a regularity property of algebraic N-point functions in momentum space. We can use the same techniques as in the case of three-point functions. Finally, we will control the canonical scaling limit in (\ref{i}) and construct pointlike localized conformal N-point functions. We present the following generalization of the conformal cluster theorem proved above (cf.\ \cite{FrJ}) to algebraic N-point functions of local observables: \medskip {\bf Theorem:} Let $(\A(I))_{I\in\KKK}$ be a conformally covariant local net on $\R$\,. Let $a_i,b_i\in\R\,,\ i=1,2,3,...\,,N$\,, and $a_i<b_i<a_{i+1}<b_{i+1}$ for $i=1,2,3,...\,,N\!-\!1$\,. Let $A_i\in\A(\,(a_i,b_i)\,)$\,, $n_i\in\N$\,, and \be P_k\,A_i\,\OM=P_k\,A_i^*\,\OM=0\,,\;k<n_i\,,\;i=1,2,3,...\,,N\,. \ee $P_k$ here denotes the projection on the subrepresentation of $U(SL(2,R))$ with conformal dimension $k$\,. We then have for each set of exponents $c_{ij}$ fulfilling the consistency conditions \be \sum_{j=1\atop j\neq i}^N c_{ij}=2n_i\,,\ c_{ij}=c_{ji}\,,\ i=1,2,3,...\,,N\,, \ee the following bound: \beam \lefteqn{ |(\,\OM,\,\left(\prod_{1\leq i\leq N}A_i\right)\OM\,)|}\nn\\ &\leq& \left(\prod_{1\leq i<j\leq N}\left|\frac{(a_i-b_i)+(a_j-b_j)}{(a_j-a_i)+(b_j-b_i)}\right|^{c_{ij}}\right) \;\prod_{1\leq i\leq N}\|A_i\|\,. \eeam If we additionally assume \be a_1-b_1=a_2-b_2=...=a_N-b_N\,, \ee we can introduce conformal cross ratios and get \beam \lefteqn{ |(\,\OM,\,\left(\prod_{1\leq i\leq N}A_i\right)\OM\,)|}\nn\\ &\leq& \left(\prod_{1\leq i<j\leq N}\left(\frac{(a_i-b_i)\,(a_j-b_j)}{(a_i-a_j)\,(b_i-b_j)}\right)^{c_{ij}/2}\right) \;\prod_{1\leq i\leq N}\|A_i\|\,. \eeam \medskip {\bf Proof:} If we pay attention to the obvious modifications needed for the additional variables, we can use in this proof the assumptions, the notation, and the line of argument introduced in the proof of the cluster theorem in the case of three-point functions. \\ We choose an arbitrary set of exponents $c_{ij}$ fulfilling the consistency conditions \be \sum_{j=1\atop j\neq i}^N c_{ij}=2n_i\,,\ c_{ij}=c_{ji}\,,\ i=1,2,3,...\,,N\,. \ee Let $R>0$\,. We consider the generator ${\rm \bf H}_R$ of the following one-parameter subgroup of $SL(2,\R)$\,: \be g_t\,:\,x\longmapsto\frac{x\, \mb{cos}\frac{t}{2}+R\,\mb{sin}\frac{t}{2}} {-\frac{x}{R}\,\mb{sin}\frac{t}{2}+\mb{cos}\frac{t}{2}}\,. \ee We know that ${\rm \bf H}_R$ is within each subrepresentation of $U(SL(2,R))$ unitarily equivalent to the conformal Hamiltonian ${\rm \bf H}$. Therefore, the spectrum of $A_i\,\OM$ and $A^*_i\,\OM$ with respect to ${\rm \bf H}_R$ is bounded from below by $n_i$\,, $i=1,2,...\,,N$\,. Let $0<t^-_{ij}<t^+_{ij}<2\pi$ such that \be g_{t^-_{ij}}(b_i)=a_j \ee and \be g_{t^+_{ij}}(a_i)=b_j\,, \ee for $i,j=1,2,...\,,N$\,, $i<j$\,. We introduce \be F(z_1,...\,,z_N)\ :=\ (\,\OM,\,\left(\prod_{i=1}^N z_{p(i)}^{-{\rm \bf H}_R}\,A_{p(i)}\, z_{p(i)}^{{\rm \bf H}_R}\right)\, \OM\,) \ee in a domain of definition given by \be |z_{p(1)}|<|z_{p(2)}|<...<|z_{p(N)}| \ee with permutations $(\,p(1),p(2),...\,,p(N)\,)$ of $(\,1,2,...\,,N\,)$\,. This definition can uniquely be extended in analogy to the case of three-point functions to boundary points with $|z_{j}|=|z_{k}|$\,, $j,k=1,2,...\,,N$\,, $j\neq k$\,, if \be g_{t_k}([a_k,b_k])\cap g_{t_j}([a_j,b_j])\neq\emptyset\,, \ee thereby introducing \be t_i:=-i\log z_i\,,\ i=1,2,...\,,N\,. \ee The line of argument presented above in the case of three-point functions and developed for general Wightman functions in \cite{StW,Jos} proves that this continuation is still an analytic function. We then define \be G(z_1,...\,,z_N)\ :=\ F(z_1,...\,,z_N)\,\prod_{1\leq i<j\leq N}\,(\frac{z_i}{z_j}-z^0_{ij})^{c_{ij}/2}\,(\frac{z_j}{z_i}-z^0_{ji})^{c_{ji}/2}\,, \ee using the abbreviation \be z_{ij}^0:=e^{i(t^-_{ij}+t^+_{ij})/2}\,,\ i,j=1,2,...\,,N\,. \ee This function is constructed such that with the consistency conditions for $c_{ij}$ and with the bound on the spectrum of ${\rm \bf H}_R$ we get the following result in analogy to the cluster theorem for three-point functions: At the boundary points $z_i=0$ and $z_i=\infty$\,, $i=1,2,...\,,N$\,, the function $G$ is bounded and can therefore be analytically continued. As in the case of three-point functions, we get with the maximum principle for analytic functions further estimates on $G$\,: Iterating the well-known maximum principle argument for the single variables, one obtains \be |G(1,...\,,1)|\ \leq\ \mbox{sup}_B\ |G(z_1,...\,,z_N)|\,, \ee where $B$ denotes the set of boundary points \be B\ :=\ \{\,|z_{j}|=|z_{k}|\ |\ g_{t_k}([a_k,b_k])\,\cap\, g_{t_j}([a_j,b_j]) \neq\emptyset\,,\ j\neq k \} \ee with $t_i=-i\log z_i$\,, $i=1,2,...\,,N$\,. The supremum of the boundary values of $G$ can be calculated in full analogy to the case of the three-point functions and to the proof of the conformal cluster theorem (cf.\ \cite{FrJ}). We obtain straightforward: \beam |(\,\OM,\,\left(\prod_{1\leq i\leq N}A_i\right)\OM\,)| &\leq& \left(\prod_{1\leq i\leq N}\|A_i\|\right)\, \prod_{1\leq i<j\leq N}\,\left|\frac{\mb{sin}\frac{t^-_{ij}-t^+_{ij}}{4}}{\mb{sin} \frac{t^-_{ij}+t^+_{ij}}{4}}\right|^{c_{ij}}.\;\; \eeam This estimate converges in the limit $R\downarrow 0$ with \be \lim_{R\rightarrow\infty}R\,t^-_{ij}=2(a_j-b_i) \ee and \be \lim_{R\rightarrow\infty}R\,t^+_{ij}=2(b_j-a_i) \ee for $i,j=1,2,...\,,N$ to the first bound asserted in the theorem. If we assume \be a_1-b_1=a_2-b_2=...=a_N-b_N\,, \ee we find \be \left(\frac{t^-_{ij}-t^+_{ij}}{t^-_{ij}+t^+_{ij}} \right)^2=\frac{(a_i-b_i)\,(a_j-b_j)}{(a_i-a_j)\,(b_i-b_j)}=r_{ij}\,,\;\;\;\, i,j=1,2,...\,,N\,, \ee and get the second bound. Hence, the theorem is proven.\hfill $\Box$ \medskip For each consistent set of exponents $c_{ij}$\,, $i,j=1,2,3,...\,,N$\,, we have proved a different bound on conformal four-point functions of chiral local observables. Hence, we know that the algebraic N-point function \be (\,\OM,\,\left(\prod_{1\leq i\leq N}\,U(-x_i)\,A_i\,U(x_i)\right)\,\OM\,) \ee decreases in position space at least as fast as the set of associated pointlike N-point functions known from conventional conformal field theory. Therefore, the Fourier transform of the algebraic N-point function has to be at least as regular in momentum space as the Fourier transforms of the associated pointlike N-point functions known from conventional conformal field theory. Technically, we follow the line of argument in the case of three-point functions and use the formula \be {\cal F}(\mbox{Pol}(X)S)=\mbox{Pol}\left(\frac{\partial}{\partial Y}\right){\cal F}S \ee for arbitrary temperate distributions $S$ and polynomials Pol$(\cdot)$ with a (multi-dimensional) variable $X$ in position space and an appropriate associated differential operator $\frac{\partial}{\partial Y}$ in momentum space. $\F$ denotes the Fourier transformation from position space to momentum space. Now, we choose $S$ to be an algebraic N-point function \be (\,\OM,\,\left(\prod_{1\leq i\leq N}\,U(-x_i)\,A_i\,U(x_i)\right)\,\OM\,) \ee of local observables $A_i$ with minimal conformal dimensions $n_i\,,\ i=1,2,...\,,N\,,$ and $X$ to be a tuple of $N-1$ algebraicly independent difference variables out of $x_i-x_j$\,, $i,j=1,2,...\,,N\,.$ The estimates in the cluster theorem proved above imply, that appropriate homogeneous polynomials $\mbox{Pol}(X)$ of degree \be \mbox{deg}(\mbox{Pol})\ =\ \left(\sum_{i=1}^N n_i\right) -2N+2 \ee can be found such that the product $\mbox{Pol}(X)\,S$ is still absolutely integrable in position space. We then see that $\mbox{Pol}(\frac{\partial}{\partial Y}){\cal F}S$ is continuous and bounded in momentum space. By locality and the condition of positive energy, the Fourier transform $F$ of an arbitrary (truncated) algebraic N-point function is known to be of the form \be F(p_1,...\,,p_{N-1})\ =\ G(p_1,...\,,p_{N-1})\ \prod_{i=1}^{N-1}\,\Theta(p_i)\,, \ee where $G$ denotes a sum of restrictions of appropriate analytic functions to subsets of momentum space (cf.\ the case of three-point functions in the section above). One can now proceed in analogy to the argumentation in the case of three-point functions: In a situation with conformal covariance and minimal conformal dimensions $n_i$\,, $i=1,2,...\,,N$\,, the function $G$ can be expressed as the product of an appropriate homogeneous polynomial $P$ of degree \be \mbox{deg}(P)\ =\ \left(\sum_{i=1}^N n_i\right) -N+1 \ee and an appropriate function $H$\,, where $H$ denotes another sum of restrictions of analytic functions to subsets of momentum space. Hence, we have proved that the Fourier transform of the algebraic N-point function \be (\,\OM,\,\left(\prod_{1\leq i\leq N}\,U(-x_i)\,A_i\,U(x_i)\right)\,\OM\,)\, \ee can be written as \be F(p_1,...\,,p_{N-1})\ =\ P(p_1,...\,,p_{N-1})\ H(p_1,...\,,p_{N-1})\ \prod_{i=1}^{N-1}\,\Theta(p_i) \ee with an appropriate homogeneous function $P$ of degree \be \mbox{deg}(P)\ =\ \left(\sum_{i=1}^N n_i\right) -N+1 \ee and an appropriate continuous and bounded function $H$\,. Using this result, we can now show in full analogy to the procedure in the last section that by canonically scaling an algebraic N-point function we construct a sequence of distributions that converges to an appropriate pointlike localized N-point function of conventional conformal field theory: \beam &&\nn\\ \lefteqn{ \lim_{\lambda\downarrow 0}\l^{-\left(\sum_{1\leq i\leq N}n_i\right)} \ (\,\OM,\,\left(\prod_{1\leq i\leq N}\,U(- \frac{x_i}{\lambda})\,A_i\,U( \frac{x_i}{\lambda})\right)\,\OM\,)} \nn\\ &&\nn\\ &=&\lim_{\lambda\downarrow 0}\l^{-\left(\sum_{1\leq i\leq N}n_i\right)} \ {\cal F}_{p_i\rightarrow x_i-x_i+1}\ F(\l p_1,...\,,\l p_{N-1})\,\l^{N-1}\,\prod_{1\leq i\leq N-1}dp_i\nn\\ &&\nn\\ &=&\lim_{\lambda\downarrow 0} \ {\cal F}_{p_i\rightarrow x_i-x_i+1}\ P(p_1,...\,,p_{N-1})\ H(\l p_1,...\,,\l p_{N-1})\ \prod_{1\leq i\leq N-1}\ \Theta(p_i)\ dp_i\nn\\ &&\nn\\ &=&\left(\prod_{1\leq i<j\leq N}\frac{1}{(x_j-x_i+i\e)^{c_{ij}}}\right)\; f(r_{t_1u_1}^{v_1s_1},...\,,r_{t_{N-3}u_{N-3}}^{v_{N-3}s_{N-3}})\,.\\ &&\nn \eeam Again, $f(\cdot,...\,,\cdot)$ denotes an appropriate function depending on $N\!-\!3$ algebraicly independent conformal cross ratios \be r_{tu}^{vs}:=\frac{(x_v-x_s)}{(x_v-x_t)}\frac{(x_t-x_u)}{(x_s-x_u)}\,. \ee The exponents $c_{ij}$ must fulfill the consistency conditions \be \sum_{j=1\atop j\neq i}^N c_{ij}=2n_i\,,\ c_{ij}=c_{ji}\,,\ 1\leq i\leq N\,, \ee which do not fully determine the exponents. Hence, the general form of the pointlike localized conformal correlation functions constructed from algebraic quantum field theory has been determined to be exactly the general form of the N-point functions known from conventional conformal field theory. In both approaches conformal covariance does not fully determine the form of N-point functions for $N>4$\,. \subsection{Wightman Axioms and Reconstruction Theorem} The most common axiomatic system for pointlike localized quantum fields is the formulation of Wightman axioms given in \cite{StW} and \cite{Jos}. (If braid group statistics has to be considered and the Bose-Fermi alternative does not hold in general, the classical formulation of \cite{StW} and \cite{Jos} has to be modified for the charged case by introducing the axiom of weak locality instead of locality \cite{FRS1,FRS2}.) The construction of pointlike localized correlation functions in this paper uses sequences of algebraic correlation functions of local observables. The algebraic correlation functions obviously fulfill positive definiteness, conformal covariance, locality, and the spectrum condition. Hence, if the sequences converge, the set of pointlike limits of algebraic correlation functions fulfills the Wightman axioms (see \cite{StW}) by construction. By the reconstruction theorem in \cite{StW} and \cite{Jos}, the existence of Wightman fields associated with the Wightman functions is guaranteed and this Wightman field theory is unique up to unitary equivalence. We do not know at the moment whether the Wightman fields can be identified with the pointlike localized field operators constructed in \cite{FrJ} from the Haag-Kastler theory. We do not know either whether the Wightman fields are affiliated to the associated von Neumann algebras of local observables and how the Haag-Kastler net we have been starting from can be reconstructed from the Wightman fields. Possibly, the Wightman fields cannot even be realized in the same Hilbert space as the Haag-Kastler net of local observables. We do know, however, that the Wightman theory associated with the Haag-Kastler theory is non-trivial: The two-point functions of this Wightman fields are, by construction, identical with the two-point functions of the pointlike localized field operators constructed in \cite{FrJ}. And we have already proved that those pointlike field vectors can be chosen to be non-vanishing and that the vacuum vector is cyclic for a set of all field operators localized in an arbitrary interval. It shall be pointed out again that those pointlike fields constructed in \cite{FrJ,Joe3} could not be proved to fulfill the Wightman axioms, since we were not able to find a domain of definition that is stable under the action of the field operators. To summarize this paper, we state that starting from a chiral conformal Haag-Kastler theory we have found a canonical construction of non-trivial Wightman fields. The reconstruction of the original net of von Neumann algebras of local observables from the Wightman fields could not explicitly be presented, since we do not know whether the Wightman fields can be realized in the same Hilbert space as the Haag-Kastler net. Actually, Borchers and Yngvason \cite{BoY} have investigated similar situations and have shown that such problems can occur in quantum field theory. In \cite{BoY} the question is discussed under which conditions a Haag-Kastler net can be associated with a Wightman theory. The condition for the locality of the associated algebra net turned out be a property of the Wightman fields called ``central positivity". Central positivity is fulfilled for Haag-Kastler nets and is stable under pointlike limits \cite{BoY}. Hence, the Wightman fields constructed in this thesis fulfill central positivity. The possibility, however, that the local net has to be defined in an enlarged Hilbert space could not be ruled out in general by \cite{BoY}. Furthermore, it has been proved in \cite{BoY} that Wightman fields fulfilling generalized H-bounds (cf.\ \cite{DSW}) have associated local nets of von Neumann algebras that can be defined in the same Hilbert space. The closures of the Wightman field operators are then affiliated to the associated local algebras. We could not prove generalized H-bounds for the Wightman fields constructed in this thesis. Actually, we suppose that the criterion of generalized H-bounds is too strict for general conformal -- and therefore massless -- quantum field theories. (Generalized) H-bounds have been proved, however, for massive theories, i.e.\ for models in quantum field theory with massive particles (cf.\ also \cite{DrF,FrH,Sum,Buc1}). \paragraph{Acknowledgements}\nichts\\ This paper is one part of the author's dissertation. We would like to thank Prof.\,Dr.\,K.\,Fre\-den\-ha\-gen for his confidence, constant encouragement, and the numerous inspiring discussions over the whole period of the work. I am indebted to him for many important insights I received by his guidance. His cooperation was crucial and fruitful for this work. The financial support given by the Friedrich-Ebert-Stiftung is gratefully acknowledged. {\small
proofpile-arXiv_065-439
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} A goal of the future heavy ion collision experiments at the relativistic heavy ion collider (RHIC) at Brookhaven and at the large hadron collider (LHC) at CERN is to find the quark-gluon plasma. The primary aim is of course to show that quarks and gluons can indeed be freed from their hadronic ``prison'' and exist as individual entities in a hot plasma. Once this is realized, one can then turn to the diverse physics of such a new state of matter. One of these is the relation of the various thermodynamic variables to each other or in other words, the equation of state \cite{shury1}. In order to probe this in experiments, an equilibrated quark-gluon plasma is required. In this work, we look at how far can one expect to have such a plasma in equilibrium. Because of the importance of this question, various different approaches have already been taken to address this issue. In particular, Shuryak \cite{shury2} argued that equilibration of the plasma proceeds via two stages in the ``hot gluon scenario''. First the equilibration of the gluons and then that of the quarks follows with a certain time delay. Thermal equilibration is quite short for gluon $\le$ 1 fm with high initial temperature of 440 MeV at LHC and 340 MeV at RHIC. However, these estimates are based on thermal reaction rates for large and small angle scatterings and on the assumption that one scattering is sufficient to achieve isotropy of momentum distribution. As has been shown in \cite{heis&wang1} using a family of different power behaviours for the time-dependence of the collision time, the assumption of one scattering is sufficient is a serious underestimate. With a larger number of scatterings, using the same arguments as in \cite{shury2}, the initial temperature will be lowered and the thermalization time will be increased. Also, we argue that estimates based on using the scattering rate alone is incorrect, since in a medium, one must consider the difference of the scattering going forward and backward both weighed with suitable factors of particle distribution functions. Hence, the process with the largest cross-section is not necessarily the more important. However, we will show the two-stage equilibration scenario or in other words, gluons equilibrate much faster than quarks and antiquarks. The other approach is the semi-classical parton cascade model (PCM) \cite{geig&mull,geig,geig&kap}, which is based on solving a set of relativistic transport equations in full six-dimensional phase space using perturbative QCD calculation for the interactions, predicts an equilibration time of 2.4 fm/c for Au+Au collision at 200 GeV/nucleon. This approach, which uses a spatial and momentum distribution obtained from the measured nuclear structure functions for the partons as initial state, is very complicated. Due to the finite size of the colliding nuclei, it is hard to clearly identify thermalization in terms of the expected time-dependent behaviours of the various collective variables \cite{geig}. But by fitting the total particle rapidity and transverse momentum distributions of the defined central volume, roughly identical temperatures are obtained \cite{geig} and hence the claim of thermalization. However, in terms of the same distributions of the individual parton components, this becomes less obvious to be the case \cite{geig&kap}. As was stated in \cite{geig&kap}, the momentum distributions are not perfect exponentials and therefore there is no complete thermalization in any case. We will look at this problem of equilibration using a much simpler approach which is based on the Boltzmann equation and the relaxation time approximation for the collision terms. Initially used by Baym \cite{baym} to study thermal equilibration and has subsequently been used in the study of various related problems \cite{gavin,kaj&mat,heis&wang2,wong}. The conclusion of these works is, in general, if the collision time $\theta$ which enters in the relaxation approximation, grows less fast than the expansion time $\tau$, then thermal equilibration can be achieved eventually. In the case of the quark-gluon plasma, it is not sufficient to know that equilibration will be achieved eventually because the plasma has not an infinite lifetime in which to equilibrate. We would like to know how far can it equilibrate before the phase transition. To answer such a question, we will use both the relaxation time approximation and the interactions obtained from perturbative QCD for the collision terms to determine $\theta$. This approach has been used previously to study both thermal and chemical equilibration in a gluon plasma \cite{wong} where it was found that with the initial conditions obtained from HIJING results, the gluon plasma had not quite enough time to completely equilibrate. In the present case of a quark and gluon parton plasma, quarks and gluons are treated as different particle species rather than as generic partons and so they have different time-dependent collision times. As a result, they approach equilibrium at different rates and towards different target temperatures. The latters will converge only at large times. It follows that the system can only equilibrate as one single system at large times. This lends support to the two-stage equilibration scenario \cite{shury2}. In an expanding system, particles are not in equilibrium early on because interactions are not fast enough to maintain this so they are most likely to start off free streaming in the beam direction \cite{heis&wang2,dan&gyul}. Thermalization will be seen as the gradual reduction of this free streaming effect as interactions gain pace and momentum transfer processes are put into action to bring the particle momenta into an isotropic distribution. The present approach takes into account of these effects. As in the previous work \cite{wong}, isotropic momentaneously thermalized initial conditions are used at both RHIC and LHC energies. These are obtained from HIJING results after allowing the partons to free stream until the momentum distribution becomes isotropic for the first time \cite{biro&etal1,lev&etal,wang}. From then on, interactions are turned on but the distribution becomes anisotropic again due to the tendency of the particles to continue to free stream. It is the role of interactions to reduce this and to progressively bring the distributions into the equilibrium forms. We have shown that, surprisingly, kinetic equilibration in a pure gluon plasma is driven mainly by gluon multiplication and not gluon-gluon elastic scattering. In this paper, we include quarks and antiquarks and consider the equilibration of a proper QCD plasma. We explicitly break down the equilibration process into each of its contributing elements and show which interactions are more important and hence uncover the dominant processes for equilibration. In fact, our result is {\em inelastic} interactions are most important for this purpose both for quarks and for gluons. Our paper is organized as follows. In Sect. \ref{sec:relax}, we describe the Boltzmann equations with the relaxation time approximation for two particle species. In Sect. \ref{sec:therm}, the time-dependent behaviour of the collision times, $\theta$'s, necessary for equilibration will be analysed and extracted. The particle interactions entering into the collision terms and details of their calculations will be explained in Sect. \ref{sec:cal}. Initial conditions used will be given in Sect. \ref{sec:ic} and lastly the results of the evolution of the plasma will be shown and discussed in Sect. \ref{sec:result}. We finish with a brief discussion of the differences with the results of PCM. \section{Relaxation Time Approximation for Two Particle Species} \label{sec:relax} In the absence of relativistic quantum transport theory derived from first principle of QCD \cite{heinz1,elze&etal,elze&heinz,elze,heinz2,heinz3}, we base our approach on Boltzmann equation with both the relaxation time approximation for the collision terms and the real collision terms obtained from perturbative QCD. Treating quarks and gluons on different footings, we write down the Boltzmann equations \begin{equation} {{\mbox{$\partial$} f_i} \over {\mbox{$\partial$} t}}+{\vo v}_{{\mathbf p}\; i} \!\cdot\! {{\mbox{$\partial$} f_i} \over {\mbox{$\partial$} {\mathbf r}}} = C_i ({\mathbf p},{\mathbf r},t) \end{equation} where $f_i$ is the one-particle distribution and $C_i$ stands for the collision terms and includes all the relevant interactions for particle species $i$ and $i=g, q,\bar q$. Concentrating in the central region of the collision where we assumed to be spatially homogeneous, baryon free and boost invariant in the z-direction (beam direction) so that $f_q=f_{\bar q}$ and $f_i=f_i({\mathbf p}_\perp,{\mathbf p}'_z,\tau)$ where $p'_z =\gamma (p_z-u p)$ with $\gamma=1/\sqrt {1-u^2}$ and $u=z/t$ is the boosted particle z-momentum component and $\tau =\sqrt {t^2-z^2}$ is the proper time. Following Baym \cite{baym}, the Boltzmann equation can be rewritten as \begin{equation} {{\mbox{$\partial$} f_i} \over {\mbox{$\partial$} \tau}} \Big |_{p_z \tau} =C_i(p_\perp,p_z,\tau) \label{eq:baymeq} \end{equation} in the central region. Using the relaxation time approximation \begin{equation} C_i(p_\perp,p_z,\tau)=- {{f_i(p_\perp,p_z,\tau)-f_{eq \; i}(p_\perp,p_z,\tau)} \over \theta_i(\tau)} \; \label{eq:relaxapp} \end{equation} where $f_{eq \; i}$ is the equilibrium distribution and $\theta_i$ is the collision time for species $i$, this allows us to write down a solution to \eref{eq:baymeq}. \begin{equation} f_i({\mathbf p},\tau)=f_{0\; i}(p_\perp,p_z \tau/\tau_0) e^{-x_i} +\int^{x_i}_0 dx'_i e^{x'_i-x_i} f_{eq\; i}(\sqrt{p^2_\perp+(p_z \tau/\tau')^2},T_{eq\; i}(\tau')) \; , \label{eq:baymeqsol} \end{equation} where \begin{equation} f_{0\; i}(p_\perp, p_z \tau/\tau_0) = \Big ( \mbox{\rm exp} (\sqrt {p^2_\perp + (p_z \tau/\tau_0)^2}/T_0)/l_{0\; i} \mp 1 \Big )^{-1} \; , \end{equation} is the solution to \eref{eq:baymeq} when $C=0$ which is also the distribution function at the initial isotropic time $\tau_0$, with initial fugacities $l_{0\; i}$ and temperature $T_0$. It is of such a form because of the assumption of momentaneously thermalized initial condition. The functions $x_i(\tau)$'s, given by \begin{equation} x_i(\tau)=\int^\tau_{\tau_0} d\tau'/ \theta_i(\tau') \; , \end{equation} play the same role as $\theta_i$'s in the sense that their time-dependent behaviours control thermalization. $T_{eq\; i}$, that appears in $f_{eq\; i}$, is the time-dependent momentaneous target equilibrium temperature for the $i$ particle species. The two terms of equation \eref{eq:baymeqsol} can be thought of, up to exponential factor, as the free streaming (first term) and equilibrium term (second term). Whether species $i$ equilibrates or not depends on which of the two terms dominates. In the present case of two species, the energy conservation equations are, in terms of the equilibrium ideal gas energy densities $\epsilon_{eq\; g}=a_2 \Tg^4$, $\epsilon_{eq\; q}=n_f b_2 \Tq^4$, $a_2=8 \pi^2/15$, $b_2=7\pi^2/40$ and $n_f$ is the number of quark flavours, \begin{equation} {{d \epsilon_i} \over {d \tau}}+{{\epsilon_i+p_{L\; i}} \over \tau} =-{{\epsilon_i-\epsilon_{eq\; i}} \over \theta_i} \end{equation} and \begin{equation} {{d \epsilon_{tot}} \over {d \tau}}+{{\epsilon_{tot}+p_{L\; tot}} \over \tau} =0 \; , \end{equation} where $\epsilon_{tot}=\sum_i \epsilon_i$ and $p_{L\; tot}=\sum_i p_{L\; i}$, or in other words \begin{equation} \sum_i {{\epsilon_i-\epsilon_{eq\; i}} \over \theta_i} =0 \; . \label{eq:e_cons} \end{equation} The above equation only expresses the fact that energy loss of one species must be the gain of the other. The transport equations of the different particle species are therefore coupled as they should be. The longitudinal and transverse pressures are defined as before \begin{equation} p_{L,T\; i}(\tau)=\nu_i \int \frac{d^3 \P}{(2\pi)^3} {{p_{z,x}^2} \over p} f_i (p_\perp,p_z,\tau) \; , \label{eq:pres} \end{equation} with $\nu_g=2\times 8=16$ and $\nu_q=2\times 3\times n_f=6\, n_f$, the multiplicities of gluons and quarks respectively. Here the equilibrium target temperatures $\Tg$ and $\Tq$ cannot be the same in general since, as we will see in Sect. \ref{sec:result}, $\qg \ne \qq =\theta_{\bar q}$. Therefore gluons and quarks will approach equilibrium at different rates. Note that energy conservation here {\em does not} mean \begin{equation} \epsilon_g+2 \epsilon_q = \epsilon_{eq\; g} + 2 \epsilon_{eq\; q} \label{eq:eeqe} \end{equation} since $\qg < \qq$ always, at least at small times, so gluon energy density $\epsilon_g$ will approach $\epsilon_{eq\; g}$ faster than $\epsilon_q$ approaches $\epsilon_{eq\; q}$ so the two equilibrium energy densities should not be considered to be those which can coexist at the same moment. This can only be true at large $\tau$ when $\Tg \simeq \Tq$ and $\qg \simeq \qq$. If \eref{eq:eeqe} were true, the condition for energy conservation \eref{eq:e_cons} could not hold when $\qg \neq \qq$. Since our QCD plasma is a dynamical system under one-dimensional expansion as well as particle production, the target temperatures $\Tg$ and $\Tq$ must be changing continuously and must approach each other at large times before the gluon and quark (antiquark) subsystems can merge into one system and exist at one single temperature. Likewise, we believe $\qg$ and $\qq$ should also converge to a single value at large times, unfortunately, this will take too long to happen in the evolution of our plasma although we can be sure that both $\qg$ and $\qq$ increase less fast than the expansion time $\tau$ near the end of the evolution, a condition which, as has already been stated in the introduction and we will see again in Sect. \ref{sec:therm}, is necessary for thermalization. \section{Conditions on $\qg$ and $\qq$ for Thermalization} \label{sec:therm} Before considering the evolution of the QCD plasma under real interactions, we can deduce analytically, using \eref{eq:baymeq} and \eref{eq:baymeqsol}, the conditions on the $\theta_i$'s under which the plasma will come to kinetic equilibrium. Multiplying \eref{eq:baymeqsol} by particle energy and integrating over momentum, we have the equations for the $\epsilon_i$'s. Further manipulating these gives, \begin{equation} \int^{x_i}_0 d x'_i \; e^{x'_i} \Big \{ \tau' h(\tau'/\tau) \Big (\epsilon_{eq\; i}(\tau')-\epsilon_i(\tau') \Big ) -{d \over {d x'_i}} \Big (\tau' h(\tau'/\tau) \epsilon_i(\tau') \Big ) \Big \} =0 \; , \label{eq:cond_theta} \end{equation} where \begin{equation} h(r)=\int^1_0 dy \sqrt {1-y^2 (1-r^2)} =\half \bigg (r+{\sin^{-1} {\sqrt {1-r^2}} \over {\sqrt {1-r^2}}} \bigg ) \end{equation} and $x'_i=x_i(\tau')$. Supposing as $\tau \rightarrow \infty$, $x_g \rightarrow \infty$ and $x_q \rightarrow \infty$ then the integrand in \eref{eq:cond_theta} will be weighed by the $\tau' \rightarrow \infty$ or large $x'_i$ limit. It follows that the term within braces in \eref{eq:cond_theta} must be zero at large $\tau'$ so using $h'(r)|_{r=1}=1/3$, we have \begin{equation} {{d \epsilon_i} \over {d \tau}}+{4 \over 3} {\epsilon_i \over \tau} = -{{\epsilon_i-\epsilon_{eq\; i}} \over \theta_i} \; . \label{eq:hydro_i} \end{equation} This means each species will undergo near hydrodynamic expansion at large $\tau$ modified by energy lost to or energy gained from the other species. The latter should be small at such times. Summing \eref{eq:hydro_i} over species, we obtain the energy conservation equation for a system undergoing hydrodynamic expansion \begin{equation} {{d \epsilon_{tot}} \over {d \tau}}+{4 \over 3} {\epsilon_{tot} \over \tau} = 0 \; , \label{eq:hydro} \end{equation} with $p_{L\; tot}=\epsilon_{tot}/3$. If one $\theta_i$ is such that the corresponding $x_i \rightarrow x_{i\; \infty} < \infty$ as $\tau \rightarrow \infty$ then hydrodynamic expansion does not apply to that species since we have \begin{equation} {{d (\epsilon_i \tau)} \over {d \tau}} =-{{(\epsilon_i-\epsilon_{eq\; i}) \tau} \over \theta_i} -p_{L\; i} \; , \label{eq:x->x<inf} \end{equation} where now $p_{L\; i} \neq \epsilon_i/3$, so kinetic equilibrium is not established. The r.h.s. of \eref{eq:x->x<inf} is negative if these particles are losing energy or gaining energy at a rate less than $p_{L\; i}/\tau$ at large $\tau$. Therefore $\epsilon_i \tau$ must decrease towards a non-zero asymptotic value $(\epsilon_i \tau)_\infty$, since $x_{i\; \infty} < \infty \Longrightarrow \epsilon_i \tau >0$ always, which results in a free streaming final state for these particles \begin{equation} \epsilon_i(\tau \rightarrow \infty) \sim (\epsilon_i \tau)_\infty /\tau \; . \label{eq:asymp_e} \end{equation} A similar free streaming final state will be reached if the rate of gaining energy is larger than $p_{L\; i}/\tau$ at large $\tau$. In this case, although $\epsilon_i \tau$ is increasing, the $\epsilon_j$ of the other particle species with $x_j \rightarrow \infty$ as $\tau \rightarrow \infty$ will be close to $\epsilon_{eq\; j}$ and so the energy transfer will be very small. One can deduce that as $\tau \rightarrow \infty$ \begin{equation} 1 \gg {{\epsilon_{eq\; i}-\epsilon_i} \over \theta_i} \rightarrow 0 \; > {p_{L\; i} \over \tau} \; \Longrightarrow \; {{d (\epsilon_i \tau)} \over {d \tau}} \rightarrow 0 \; , \label{eq:onlot} \end{equation} hence $\epsilon_i \tau\rightarrow (\epsilon_i \tau)_\infty$. That is $\epsilon_i \tau$ now increases towards some asymptotic value instead of decreasing towards one as in the previous case. But it ends up with a free streaming final state nevertheless. We do not consider the case where the relative rate $(\epsilon_{eq\; i}-\epsilon_i)\tau/\theta_i p_{L\; i}$ oscillates about one at large $\tau$ except to say that on the average $d (\epsilon_i \tau)/d \tau \sim 0$ and so an average free streaming final state is likely. The last possibility where $x_i \rightarrow x_{i\; \infty} <\infty$ as $\tau \rightarrow \infty$ for both particle species, \eref{eq:x->x<inf} applies to both. Barring the case of the oscillating relative rate, one particle species must lose energy and so by the above argument, a free streaming final state results. For the remaining particle species, it does not matter whether $d (\epsilon_i \tau)/d\tau$ is or is not positive at large $\tau$, these particles will also be in a free streaming final state. If the rate is negative, then the same argument that leads to \eref{eq:asymp_e} applies. If it is positive, since the species that is losing energy is approaching free streaming so the energy transfer must go to zero. Then we are back to \eref{eq:onlot}. The conclusions are therefore, depending on the time-dependent behaviours of $\qg$ and $\qq$, \begin{enumerate} \item $x_g \rightarrow \infty$ and $x_q \rightarrow \infty$ as $\tau \rightarrow \infty$ are required for the whole system to completely thermalize. \item $x_g \rightarrow \infty$ and $x_q \rightarrow x_{q\; \infty} < \infty$ or $x_q \rightarrow \infty$ and $x_g \rightarrow x_{g\; \infty} < \infty$ as $\tau \rightarrow \infty$ imply that only the species with $x_i \rightarrow \infty$ will thermalize, the other species will not equilibrate but free streams at the end. The system will end up somewhere between free streaming and hydrodynamic expansion. \item Both $x_g \rightarrow x_{g\; \infty} <\infty$ and $x_q \rightarrow x_{q\; \infty} <\infty$ as $\tau \rightarrow \infty$ then the whole system will end up in a free streaming final state. \end{enumerate} One can understand these $x_i$ behaviours in terms of $\theta_i$'s by assuming simple power $\tau$-dependence for the latters. One finds that $\theta_i$'s must all grow slower than $\tau$ for the whole system to achieve thermalization. If either one or more grow faster then a mixed or a complete free streaming final state results. \section{Particle Interactions --- Collision Terms} \label{sec:cal} To investigate the evolution of a proper QCD plasma, we consider the following simplest interactions at the tree level \begin{equation} gg \longleftrightarrow ggg \; \; , \; \; \; gg \longleftrightarrow gg \; , \label{eq:ggi} \end{equation} \begin{equation} gg \longleftrightarrow q\bar q \; \; , \; \; \; g q \longleftrightarrow g q \; \; , \; \; \; g\bar q \longleftrightarrow g\bar q \; , \label{eq:gqi} \end{equation} \begin{equation} q\bar q \longleftrightarrow q\bar q \; \; , \; \; \; qq \longleftrightarrow qq \; \; , \; \; \; \bar q \bar q \longleftrightarrow \bar q \bar q \; . \label{eq:qqi} \end{equation} As in \cite{biro&etal1,lev&etal,wang}, we include only the leading inelastic processes i.e. the first interaction of \eref{eq:ggi} and \eref{eq:gqi}\footnote{The first one of \eref{eq:qqi} could also be inelastic but here we give the same chemical potential to all the fermions so we do not consider quark-antiquark annihilations into different flavours as inelastic for our purpose.}. We will return to this point later on in Sect. \ref{sec:result}. In the solutions \eref{eq:baymeqsol} to the Boltzmann equations \eref{eq:baymeq}, there are two time-dependent unknown parameters $\theta_i$ and $T_{eq\;i}$ for each species which very much control the particle distributions. To determine them, we need two equations each for gluons and for quarks. In order to show the relative importance of the various interactions \etrref{eq:ggi}{eq:gqi}{eq:qqi} in equilibration, we find these time-dependent parameters by constructing equations from the rates of energy density transfer between quarks (antiquarks) and gluons and the collision entropy density rates. From \etrref{eq:baymeq}{eq:relaxapp}{eq:baymeqsol}, the energy density transfer rates are \begin{equation} {{d \epsilon_i} \over {d \tau}}+{{\epsilon_i + p_{L\; i}} \over \tau} = -{{\epsilon_i-\epsilon_{eq\; i}} \over \theta_i} = \nu_i \int \frac{d^3 \P}{(2\pi)^3} \; p \; C_i (p_\perp,p_z,\tau) = {\cal E}_i \; , \label{eq:e_trans} \end{equation} where ${\cal E}_i$ is the energy gain or loss of species $i$ per unit time per unit volume. As stated in Sect. \ref{sec:relax}, ${\cal E}_i$'s must obey $\sum_i {\cal E}_i =0$ for energy conservation. The other equations, the collision entropy rates can be deduced from the explicit expression of the entropy density in terms of particle distribution function \cite{groot} \begin{equation} s_i(\tau)=-\nu_i \int \frac{d^3 \P}{(2\pi)^3} \Big \{f_i({\mathbf p},\tau) \ln f_i({\mathbf p},\tau) \mp (1 \pm f_i({\mathbf p},\tau)) \ln (1 \pm f_i({\mathbf p},\tau)) \Big \} \; , \end{equation} where the different signs are for bosons and fermions respectively. They are, using again \etrref{eq:baymeq}{eq:relaxapp}{eq:baymeqsol}, \begin{eqnarray} \Big ( {{d s_i} \over {d \tau}} \Big )_{coll} \fx & = & \fx -\nu_i \int \frac{d^3 \P}{(2\pi)^3} \Big ( {{\mbox{$\partial$} f_i} \over {\mbox{$\partial$} \tau}} \Big )_{coll} \ln \Big ({{f_i} \over {1 \pm f_i}} \Big ) \\ \fx & = & \fx -\nu_i \int \frac{d^3 \P}{(2\pi)^3} \; C_i (p_\perp,p_z,\tau) \ln \Big ({{f_i} \over {1 \pm f_i}} \Big ) \label{eq:s_rate1} \\ \fx & = & \nu_i \int \frac{d^3 \P}{(2\pi)^3} {{f_i -f_{eq\; i}} \over \theta_i} \ln \Big ({{f_i} \over {1 \pm f_i}} \Big ) \; . \label{eq:s_rate2} \end{eqnarray} By using the explicit expression for the collision terms $C_i$'s constructed from the interactions \etrref{eq:ggi}{eq:gqi}{eq:qqi} within perturbative QCD, \etrref{eq:e_trans}{eq:s_rate1}{eq:s_rate2} allow us to solve for $\theta_i$'s and $T_{eq\; i}$'s. The gluon multiplication contribution to $C_g$ is constructed from the infrared regularized Bertsch and Gunion formula \cite{bert&gun} for the amplitude with partial incorporation of Landau-Pomeranchuk-Migdal suppression (LPM) for gluon emission and absorption \cite{biro&etal1,gyul&wang,gyul&etal,baier&etal} as in the previous work \cite{wong}. The explicit form of the gluon multiplication collision term and a discussion of the problem regarding how to incorporate the LPM effect correctly can be found there also. The remaining binary interaction contributions to $C_i$ for particle $1$ is, as usual, given by \begin{eqnarray} C_{i\; 1}^{binary} \fx & = & \fx - \sum_{{\cal P}_i} {{S_{{\cal P}_i} \nu_2} \over {2 p_1^0}} \prod^4_{j=2} {{d^3 {\mathbf p}_j} \over {(2\pi)^3 2 p_j^0}} (2\pi)^4 \delta^4(p_1+p_2-p_3-p_4) |{\cal M}^{{\cal P}_i}_{1+2 \rightarrow 3+4}|^2 \nonum & & \times [f_1 f_2(1 \pm f_3)(1 \pm f_4) -f_3 f_4 (1 \pm f_1)(1 \pm f_2)] \end{eqnarray} where the ${\cal P}_i$ runs over all the binary processes in \etrref{eq:ggi}{eq:gqi}{eq:qqi} which involve species $i$, $|{\cal M}^{{\cal P}_i}|^2$ is the sum over final states and averaged over initial state squared matrix element, $S_{{\cal P}_i}$ is a symmetry factor for any identical particles in the final states for the process ${\cal P}_i$ and $\nu_2$ is the multiplicity of particle 2. We take $|{\cal M}^{{\cal P}_i}|^2 \; $'s from \cite{cut&siv} and infrared regularized them using either the Debye mass $m_D^2$ for gluons or the quark medium mass $m_q^2$ for quarks to cut off any infrared divergence. These masses are now time-dependent quantities in a non-equilibrium environment. With non-isotropic momentum distribution, both the Debye mass \cite{biro&etal2,esk&etal} and the gluon medium mass, $m^2_g$, are directional dependent. This is, however, not the case for the quark medium mass, $m^2_q$, which remains directional independent as in equilibrium. The directional dependence arises out of the cancellations between identical type of distribution functions similar to those one finds in the derivation of hard thermal loops \cite{bra&pis,fre&tay}. To keep things simple, we removed the directional dependence from $m_D^2$ and use, for SU(N=3), to leading order in $\alpha_s$, \begin{equation} m^2_D (\tau)=-8 \pi \alpha_s \int \frac{d^3 \P}{(2\pi)^3} {\mbox{$\partial$} \over {\mbox{$\partial$} |{\mathbf p}|}} \Big (N \, f_g +n_f \, f_q \Big ) \; . \end{equation} For the quark medium mass, to the same order, we use \begin{equation} m_q^2 (\tau)= 4\pi \alpha_s \; \Big ({{N^2-1} \over {2\, N}}\Big ) \int \frac{d^3 \P}{(2\pi)^3} {1 \over {|{\mathbf p}|}} \; \big (f_g +f_q \big ) \; , \end{equation} which is just the equilibrium expression but with non-equilibrium distribution functions. With these masses, we regularize the squared matrix elements by hand and inserting the masses as follows. \begin{eqnarray} |{\cal M}_{gg \rightarrow gg}|^2 \fx &=& \fx {{9\; g^2} \over 2} \; \bigg ( 3-{{u t} \over {(s+m_D^2)^2}} -{{u s} \over {(t-m_D^2)^2}} -{{s t} \over {(u-m_D^2)^2}} \bigg ) \\ |{\cal M}_{gg \rightarrow q \bar q}|^2 \fx &=& \fx {g^2 \over 6} \; \bigg ( {t \over {(u-m_q^2)}} +{u \over {(t-m_q^2)}} \bigg ) - {3 \over 8} \; {{u^2+t^2} \over {(s+4 m_q^2)^2}} \\ |{\cal M}_{gq \rightarrow gq}|^2 = |{\cal M}_{g \bar q \rightarrow g \bar q}|^2 \fx &=& \fx g^2 \bigg ( 1- {{2 u s} \over {(t-m_D^2)^2}} -{4 \over 9} \; \bigg ( {u \over {(s+m_q^2)}} +{s \over {(u-m_q^2)}} \bigg ) \bigg ) \\ |{\cal M}_{qq \rightarrow qq}|^2 = |{\cal M}_{\bar q \bar q \rightarrow \bar q \bar q}|^2 \fx &=& \fx {{2\; g^2} \over 9} \; \bigg ( {{2(s^2+t^2)} \over {(u-m_D^2)^2}} + \delta_{12} \; {{2(u^2+s^2)} \over {(t-m_D^2)^2}} \nonum \fx & & \fx \; \; \; \; \; - \delta_{12} \; {4 \over 3} \; {s^2 \over {(t-m_D^2)(u-m_D^2)}} \bigg ) \\ |{\cal M}_{q \bar q \rightarrow q \bar q}|^2 \fx &=& \fx {{2\; g^2} \over 9} \; \bigg ( \delta_{13} \delta_{24} {{2(s^2+t^2)} \over {(u-m_D^2)^2}} + \delta_{12} \delta_{34} \; {{2(t^2+u^2)} \over {(s+4 m_q^2)^2}} \nonum \fx & & \fx \; \; \; \; \; - \delta_{12} \delta_{13} \delta_{34} \; {4 \over 3} \; {t^2 \over {(u-m_D^2)(s+4 m_q^2)}} \bigg ) \end{eqnarray} where the $\delta_{ij}$ signifies that the $i$ and $j$ quark or antiquark must be of the same flavour. This regularization amounts to screening spacelike and timelike infrared gluons by $m^2_D$ and $4 m^2_q$, respectively and infrared quarks by $m^2_q$. We stress that this regularization is done in a very simple manner and with the right order of magnitude for the cutoffs. Its aim is to get some estimates to the collision rates without involving too much with the exact and necessarily complicated momentum dependent form of the true infrared screening self-energies in an out-of-equilibrium plasma when their infrared screening effects should be in action. They should be the extension of the 2-point gluon and quark hard thermal loops \cite{bra&pis,fre&tay,weld,klim,tay&wong} to a non-thermalized environment. We should mention here that the choice of the pair of equations for solving the two time-dependent unknowns $\theta_i$ and $T_{eq\; i}$ for each particle species is not unique. One can equally use, for example, the rate equations for the particle number density instead of the collision entropy density. With these other choices, the values of the different quantities are shifted somewhat due to the way that the initial conditions are extracted but there is no qualitative different in the result. Our present choice has the distinct advantage that we can explicitly compare the different processes using the collision entropy density rates. This will become clear when we show the results in Sect. \ref{sec:result}. \section{Initial Conditions} \label{sec:ic} To start the evolution, we use the same initial conditions for the gluon plasma as before \cite{wong} based on HIJING result for Au+Au collision. The initial conditions for the quarks (antiquarks) are obtained by taking a ratio of $0.14$ for the number of initial quark (antiquark) to the initial total number of partons as done in \cite{biro&etal1,lev&etal,wang}. The initial conditions are shown in Table 1. One sees that the initial quark collision times are long compared to those of the gluons both at RHIC and LHC. Especially at RHIC, the quark collision time is exceedingly long and so these particles are essentially free streaming initially. Taking these numbers as guides to how fast each particle species is going to equilibrate, we can be sure already of a two-stage equilibration scenario \cite{shury2}. \begin{center} \begin{tabular}{|c|c|c|} \hline\hline \multicolumn{3}{|c|}{Initial Conditions} \\ \hline \emph{\ } & \ \ \emph{RHIC}\ \ & \ \ \emph{LHC}\ \ \\ \hline $\tau_0$ (fm/c) & 0.70 & 0.50 \\ $T_0$ (GeV) & 0.50 & 0.74 \\ $\epsilon_{0\; g}$ (GeV/$\mbox{\rm fm}^3$) & 3.20 & 40.00 \\ $\epsilon_{0\; q}$ (GeV/$\mbox{\rm fm}^3$) & 0.63 & 7.83 \\ $n_{0\; g} (\mbox{\rm fm}^{-3})$ & 2.15 & 18.00 \\ $n_{0\; q} (\mbox{\rm fm}^{-3})$ & 0.42 & 3.53 \\ $l_{0\; g}$ & 0.08 & 0.21 \\ $l_{0\; q}$ & 0.017 & 0.044 \\ $\theta_{0\; g}$ (fm/c) & 2.18 & 0.73 \\ $\theta_{0\; q}$ (fm/c) & 239.72 & 30.92 \\ \hline\hline \end{tabular} \end{center} \begin{center} TABLE 1. Initial conditions for the evolution of a QCD plasma created in Au+Au collision at RHIC and at LHC \end{center} \vspace{0.5cm} Using the standard initial picture of heavy ion collisions as before, our evolution is started when the momentum distribution in the central region of the collision becomes, for the first time, isotropic due to longitudinal cooling. The subsequent development is determined by the interactions \etrref{eq:ggi}{eq:gqi}{eq:qqi}. In the case of a pure gluon plasma \cite{wong}, it is clear that interactions bring the system towards equilibrium and not towards some free streaming final state which is a possible alternative as can be inferred from the analysis in Sect. \ref{sec:therm}. That is the interactions dominate over the expansion. In the present situation, we will see that the same can certainly be said for the gluons and for the quarks at LHC but at RHIC, it is less clear for the latters. The equilibration time for quarks is at least several times longer than that of the gluons. Details for the procedure of the computation can be found in \cite{wong}. The values for the numerical parameters are the same and in addition, we use $n_f=2.5$ to take into account of the reduced phase space of strange quark. All time integrations are discretized and the rates are obtained at each time step necessary for forming the two pairs of equations \etrref{eq:e_trans}{eq:s_rate1}{eq:s_rate2}. One then solves the two equilibrium temperatures $\Tg$ and $\Tq$ from two 4th degree polynomials, one for each of the temperatures. From these solutions, $\qg$ and $\qq$ are obtained and everything is then fed back into the equations for the next time step. \section{Equilibration of the QCD Plasma} \label{sec:result} We show the results of our computation in this section. They show clearly the collision times $\qg$ and $\qq$ hold the keys to equilibration as have been analysed in Sect. \ref{sec:therm}. We will see shortly that as a result of the disparity between their magnitudes at finite values of $\tau$, the equilibration of quarks and antiquarks lags behind that of the gluons both chemically and kinetically. We will also identify the dominant processes responsible for equilibration. They are {\em not} the commonly assumed elastic scattering processes as already mentioned in the introduction. When dealing with two particle species, one has several choices as to when should the evolution be stopped. We choose to do this when both the quark and the gluon temperature estimates drop to 200 MeV. For gluons, this estimate is obtained by the near equilibrium energy and number density expression \begin{equation} \epsilon_g = a_2 \, l_g \, T^4_g \mbox{\hskip 1cm and \hskip 1cm} n_g = a_1 \, l_g \, T^3_g \; , \end{equation} which are valid when the fugacity $l_g$ is near $1.0$ i.e. when the distribution functions can be approximated by $f_g({\mathbf p},\l_g,\tau) = l_g f_g({\mathbf p},\l_g=1,\tau)$. For quarks and antiquarks, we cannot do the same as $l_q$ has not time to rise above $0.5$ so instead, the temperature is estimated from the same quantities in kinetic equilibrium but at small values of $l_q$ \begin{equation} \epsilon_q = 3 \, \nu_q \, l_q \, T^4_q / \pi^2 \mbox{\hskip 1cm and \hskip 1cm} n_q = \nu_q \, l_q \, T^3_q / \pi^2 \; . \end{equation} \begin{figure} \centerline{ \hbox{ {\psfig{figure=fig1a.ps,width=3.4in}} \ {\psfig{figure=fig1b.ps,width=3.4in}} }} \caption{The time-dependence of the estimated temperatures for quarks and for gluons and their fugacities at (a) LHC and (b) RHIC. The solid lines are the estimated temperatures $T_g$ (thick line) and $T_q$. The dashed lines are the fugacities $l_g$ (thick line) and $l_q$. Gluon chemical equilibration is much faster than that of the quarks. The curves are stopped when all the temperature estimates drop to 200 MeV. The vertical line indicates when the gluon temperature reaches this value.} \label{gr:T&l} \end{figure} \noindent These estimates are plotted in \fref{gr:T&l}. The vertical line marks the point when the gluon temperature estimate (thick solid line) drops to $200$ MeV. At this point, $\tau \sim 6.25$ fm/c, the fugacity (thick dashed line) is $l_g \sim 0.935$ at LHC and is $l_g \sim 0.487$ at $\tau \sim 2.85$ fm/c at RHIC. On the same plots, the quark temperature (solid line) drops at a slower rate and the fermionic fugacity (dashed line) is also increasing much slower given the less favourable initial conditions and initially much slower quark-antiquark pair creation than gluon multiplication rate. In the end, the fermions are not too well chemically equilibrated and in fact, are still quite far away from $1.0$. This is especially bad at RHIC. We note that comparing to \cite{biro&etal1,lev&etal,wang}, in our case, gluons chemically equilibrate faster but quarks are slower. Unlike chemical equilibration, kinetic equilibration has no simple indicators like the fugacities that can allow itself to be simply quantified. One has to, instead, use the anisotropy of momentum distribution as well as various reaction rates to get an idea of the degree of kinetic equilibration. The former can be deduced from the ratios of the longitudinal pressure and a third of the energy density to the transverse pressure, $p_L/p_T$ and $\epsilon/3 p_T$ respectively. Whereas from the elastic scattering rates, one can deduce roughly how close the distribution functions are to their equilibrium forms by virtue of the fact that in local kinetic equilibrium, these rates are zero. The pressure ratios $p_L/p_T$ (solid line) and $\epsilon/3 p_T$ (dashed line) are plotted in \fref{gr:press} (a) and (a') for gluons, (b) and (b') for quarks and (c) and (c') for the total sum. These ratios are indeed approaching $1.0$, the expected value after thermalization, but at different rates. Gluons are clearly equilibrating much faster than quarks which proceed rather slowly. \begin{figure} \centerline{ \hbox{ {\psfig{figure=fig2.ps,width=3in}} \ \ {\psfig{figure=fig2da.ps,width=3in}} }} \caption{The ratios of the longitudinal pressure (solid line) and a third of the energy density (dashed line) to the transverse pressure, $p_L/p_T$ and $\epsilon/3 p_T$ respectively for (a) gluons, (b) quarks and (c) the total sum at LHC. Graphs (a'), (b') and (c') are the same at RHIC.} \label{gr:press} \end{figure} To show that these behaviours, although slow, are indeed the signs of equilibration and that the plasma is not approaching some free streaming final states, we can work out what their behaviours should be in the latter case by taking the extreme and let $\theta_i \rightarrow \infty$. From \eref{eq:pres}, as $\tau \rightarrow \infty$, \vspace{2ex} \hbox{\hspace{4.9cm} \raisebox{-2.5ex}{\vbox{ \hbox{ \( p_L \rightarrow \pi \, \tau_0^3 \, \epsilon_0 /4 \, \tau^3 \) } \hbox{ \( p_T \rightarrow \pi \, \tau_0 \, \epsilon_0 /8 \, \tau \) } \hbox{ \( \; \; \epsilon \: \rightarrow \pi \, \tau_0 \, \epsilon_0 /4 \, \tau \) } }} \( \bigg \} \; \; \Longrightarrow \; \; \bigg \{ \) \raisebox{-1ex}{\vbox{ \hbox{ \( p_L/p_T \rightarrow 2 \, \tau^2_0/ \tau^2 \rightarrow 0 \) } \hbox{ \( \epsilon/3 \: p_T \rightarrow 2/3 \) } }} \vbox{\parbox{4cm}{ \begin{equation} \; \end{equation} }} } \vspace{1ex} \noindent where $\epsilon_0$ is the initial energy density and the above ratios are valid for both quarks and gluons in this extreme. Therefore in the free streaming case, the first ratio should approach zero and the second should approach 2/3. These are clearly not what we see in our plots. \begin{figure} \centerline{ \hbox{ {\psfig{figure=fig3.ps,width=3in}} \ \ {\psfig{figure=fig3da.ps,width=3in}} }} \caption{The scaled products of the collective variables (a) energy density, (b) number density and (c) entropy density and their expected inverse time-dependence in equilibrium $\tau^{4/3}$, $\tau$ and $\tau$ respectively at LHC. Graphs (a'), (b') and (c') are the same at RHIC. The solid and dashed lines are for gluons and quarks respectively. The thick solid line in (c) and (c') is the scaled product of the total entropy density and $\tau$.} \label{gr:scaled} \end{figure} To best get an idea of how close the distribution functions are to the equilibrium forms, the $gg$ and $qq$ or $\bar q\bar q$ elastic scattering processes are ideal for this. These are shown in \fref{gr:s_g-lhc} and \fref{gr:s_g-rhic} (b) for gluon and \fref{gr:s_q-lhc} and \fref{gr:s_q-rhic} (c) for quark. Note that the peaks of these collision entropy rates coincide with the corresponding mininum points of the pressure ratios. As expected, the rates maximize at maximum anisotropy in momentum distribution. They all rise rapidly from zero at $\tau_0$ when the interactions are turned on. The subsequent return to zero or the approach of the distribution functions to their equilibrium forms are, however, much less rapid. They only do so progressively as can be deduced already from the pressure ratio plots. Having shown chemical and kinetic equilibrations separately, we present now the actual approach of the collective variables towards the equilibrium values. Since we are more interested in the behaviour of their time-dependence than their absolute magnitudes, we multiplied them by their expected time-dependence and scaled these by taking a guess at the corresponding asymptotic values from the tendency of the curves. The results are plotted in \fref{gr:scaled}. They are $\epsilon_i \tau^{4/3}/\epsilon_{s\, i} \tau_{s\, i}^{4/3}$, $n_i \tau/n_{s\, i} \tau_{s\, i}$ and $s_i \tau/s_{s\, i} \tau_{s\, i}$ in the figures (a) and (a'), (b) and (b') and (c) and (c') respectively. All these should be nearly constant with respect to time at large $\tau$. The solid lines are for gluons and the dashed ones are for quarks. They showed that the curves do behave in such a way for the eventual constant behaviour. This feature is much clearer at LHC than at RHIC which only reconfirms the previously deduced result of faster equilibration at LHC than at RHIC. Note that for gluons, the quantities are approaching the corresponding asymptotic values from above, whereas for quarks, this approach is from below. This is because of the simple reason that there is a net conversion of gluons into quark-antiquark pairs via $gg \longleftrightarrow q\bar q$. The corresponding collision entropy density rate is negative as shown in \fref{gr:s_g-lhc} and \fref{gr:s_g-rhic} (c). We will see that this same interaction becomes dominant in the later part of the evolution later on when we compare the importance of the different processes. So gluons are losing energy, number and entropy to the fermions. This has to be so before the system as a whole can settle into complete equilibrium. The thick solid lines in \fref{gr:scaled} (c) and (c') show the scaled total entropy per unit area in the central region which give an idea of the state of the system as a whole. They show that although the entropy of the individual subsystem can decrease, the total value must increase in accordance with the second law of thermodynamics. \begin{figure} \centerline{ \hbox{ {\psfig{figure=fig4.ps,width=3in}} \ \ {\psfig{figure=fig4da.ps,width=3in}} }} \caption{The time-dependence of the collision time (a) for gluons $\qg$ and (b) for quarks $\qq$ at LHC. Their values are compared in (c). $\tau$ overtakes first $\qg$ and later $\qq$ also. Graphs (a'), (b') and (c') are the same at RHIC. In this case, $\tau$ only has time to overtake $\qg$ but not $\qq$.} \label{gr:relax_t} \end{figure} The figures discussed above show that the plasma is indeed approaching equilibrium and that interactions are fast enough to dominate over the Bjorken type one-dimensional scaling expansion. As we analysed in Sect. \ref{sec:therm}, thermalization is governed by the $\theta_i$'s. How fast this will proceed depends on their magnitudes and what is the actual final state depends on their time-dependent behaviours. For thermalization, the $\theta_i$'s must behave in such a way such that $x_i \rightarrow \infty$ as $\tau \rightarrow \infty$. That means they must grow less fast than $\tau$. In \fref{gr:relax_t}, we show these $\theta_i$'s as a function of $\tau$. Initially, $\theta_i >\tau$ for both quarks and gluons, and $\qq$ starts off very large (see Table 1) but drops extremely rapidly back down to within hadronic timescales. The subsequent expected increase in time \cite{baymetal1,baymetal2,heis} is sufficiently slow for $\tau$ to get past $\qg$ and $\qq$ at LHC, \fref{gr:relax_t} (a) and (b) but at RHIC, \fref{gr:relax_t} (b'), $\qq$ is still too large for $\tau$ to overtake it before the temperature reaches $200$ MeV. Nevertheless, the $\tau$-dependence is slow enough that $x_i$ should go to infinity as $\tau \rightarrow \infty$. We have mentioned in Sect. \ref{sec:relax}, for the system to equilibrate as one, the target equilibrium temperatures $\Tg$ and $\Tq$ and also $\qg$ and $\qq$ must approach each other at large $\tau$. We strongly suspect that the convergence of the temperatures will proceed in an oscillating fashion where the two curves intersect each other several times before the final convergence at very large $\tau$. We can see this in \fref{gr:T_eq} (a) and (b). At LHC, the initial condition is more favourable for equilibration and so $\Tg$ intersects $\Tq$ twice already. This is not so at RHIC. In fact, all indications point to the fact that a plasma created at LHC will equilibrate better than one created at RHIC. By letting the plasma to continue its evolution and ignoring the deconfinement phase transition, we have seen that the collective variables like the gluon and quark energy densities, gluon entropy density etc. do show tendency to pass from below to above or vice versa, the corresponding equilibrium target values i.e. tendency to overshoot the equilibrium values and hence oscillation. As to the convergence of $\theta_i$'s, it is not so clear in \fref{gr:relax_t} (c) and (c'), especially at RHIC in \fref{gr:relax_t} (c'). $\qq$ is much too large in comparison with $\qg$ for any clear sign of convergence within the time available. On the other hand, at LHC, although there is still a large gap between the magnitudes, there is a clear tendency that the rate of increase of $\qq$ with $\tau$ is slowing down in \fref{gr:relax_t} (b) while $\qg$ still increases at approximately the same rate. It is simply too early for the system to equilibrate as one. Even near the end, the quarks and gluons can only be considered as two linked subsystems approaching equilibrium at very different rates. Hence we have a two-stage equilibration. \begin{figure} \centerline{ \hbox{ {\psfig{figure=fig5a.ps,width=3.2in}} \ \ {\psfig{figure=fig5b.ps,width=3.2in}} }} \caption{The time development of the equilibrium target gluon (solid line) and quark (dashed line) temperatures $\Tg$ and $\Tq$ respectively at (a) LHC and (b) RHIC. They should converge in an oscillating fashion at large $\tau$ in order for the system to equilibrate as one towards a single temperature. The convergence is less good at RHIC than at LHC.} \label{gr:T_eq} \end{figure} Having shown that interactions can indeed dominate over the one-dimensional expansion of the parton gas in the central region of relativistic heavy ion collisions and hence bring the plasma into equilibrium. We can now look at the individual processes and compare their relative importance. These are the processes \etrref{eq:ggi}{eq:gqi}{eq:qqi}. We have labelled their contributions to the gluon and quark collision entropy rate $ds_g/d\tau$ and $ds_q/d\tau$ by $d s_{gi}/d\tau$, $i=1,\dots, 4$ and $d s_{qi}/d\tau$, $i=1,\dots,3$ in the order that they appear in \etrref{eq:ggi}{eq:gqi}{eq:qqi}. Processes that give the same rate due to quark-antiquark symmetry are considered as the same process. Hence $gq \longleftrightarrow gq$ and $g\bar q \longleftrightarrow g\bar q$ give identical contribution to gluon and quark collision entropy density rate as $ds_{g4}/d\tau$ and $ds_{q2}/d\tau$ respectively. Also we have combined fermion elastic scattering processes as one rate $ds_{q3}/d\tau$ for convenience. These are shown in \fref{gr:s_g-lhc}, \fref{gr:s_g-rhic}, \fref{gr:s_q-lhc} and \fref{gr:s_q-rhic}. The elastic processes have a characteristic shape, i.e. an initial rapid rise to a peak at maximum anisotropy before returning to zero progressively. The sharper the peak, the quicker the kinetic equilibration (compare \fref{gr:s_g-lhc} and \fref{gr:s_g-rhic} (b), (d) and \fref{gr:s_q-lhc} and \fref{gr:s_q-rhic} (b), (c) and \fref{gr:press}). Note the negative rate of \fref{gr:s_g-lhc} and \fref{gr:s_g-rhic} (c) which is because there are net quark-antiquark pair creations from gluon-gluon annihilations and entropy decreases with the number of gluons as already mentioned in the previous paragraphs. We compare the different processes by plotting the ratio of the magnitude of each contribution to that of gluon multiplication for gluons in \fref{gr:s_g-lhc} and \fref{gr:s_g-rhic} (e) and the ratio of each rate to that of quark-antiquark creation for quarks (antiquarks) in \fref{gr:s_q-lhc} and \fref{gr:s_q-rhic} (d). In the (e) figures, gluon multiplication clearly dominates initially at $\tau \lsim 2$ fm/c at LHC and $\tau \lsim 4$ fm/c at RHIC since all three ratios in each plot are less than 1. After these times, $q\bar q$ creation becomes dominant (thick solid line) and rises to several times larger than gluon multiplication. The $gg$ elastic scattering, on the other hand, tends to maintain a small, nearly constant ratio with gluon multiplication (solid line), which supports the claim made in \cite{wong}. That is, in a pure gluon plasma, gluon multiplication dominates over $gg$ elastic scattering in driving the plasma towards equilibrium. This remains the case even when $l_g \sim 0.93$ which shows that this dominance is not sensitive to the value of $l_g$. The remaining ratio of quark-gluon scattering to gluon multiplication continues to rise but not as rapidly as the first ratio. For quark entropy, \fref{gr:s_q-lhc} and \fref{gr:s_q-rhic} (d), both ratios of quark-gluon scattering (solid line) and fermion-fermion scatterings (dashed line) to $gg \longleftrightarrow q\bar q$ rate remain small during the time available although they are both on the rise. So for gluons, gluon multiplication dominates initially but is later overtaken by $gg\longleftrightarrow q\bar q$ which continues to dominate over other elastic processes. For quarks (antiquarks), this same process dominates during the lifetime of the plasma. These behaviours can be understood in the following way. Gluon branching dominates initially over any other processes so long as gluons are not near equilibrium. Once they approach saturation (the $l_g$ estimates slow down their approach towards $1.0$ in \fref{gr:T&l} (a) and (b) at about the times mentioned above), gluon-gluon annihilation to quark-antiquark takes over as the dominant one because the fermions are still far from full equilibration. Because of the latter reason, the other ratios involving quark or antiquark to gluon branching continue to rise. \begin{figure} \centerline{ {\psfig{figure=fig6.ps,width=6.7in}} } \caption{ The time development of the different contributions to the total gluon collision entropy density rate at LHC. They are (a) $gg\longleftrightarrow ggg$, (b) $gg\longleftrightarrow gg$, (c) $gg\longleftrightarrow q\bar q$ and (d) $gq\longleftrightarrow gq$ or $g\bar q\longleftrightarrow g\bar q$. The curves of the elastic scattering processes in (b) and (d) have typical peaks at maximum anisotropy in momentum distributions. The ratios of the contribution (b) (thick line), (c) (solid line) and (d) (dashed line) to that of (a) are plotted in (e). This shows that first gluon multiplication dominates initially but is later overtaken by gluon annihilations into quark-antiquark pairs.} \label{gr:s_g-lhc} \end{figure} So contrary to common assumption, inelastic processes are dominant in equilibration. This should have consequences in the perturbative calculations of transport coefficients or relaxation times \cite{baymetal1,baymetal2,heis} of system that are not subjected to external forces. These calculations are based essentially, up to the present, on elastic binary interactions. As we have seen, they are not the dominant processes in equilibration. To the surprising result of gluon multiplication dominates over elastic gluon-gluon scattering, we provide the following explanation. If one only looks at the scattering cross-sections, it is indeed true that gluon-gluon scattering has a larger value and gluon multiplication processes are down by $\alpha_s$ for each extra gluon produced. The $(n-2)$ extra gluon production cross-section can be expressed in terms of the elastic scattering cross-section as \cite{gold&rosen,shury&xion}, in the double logarithmic approximation, \begin{equation} \sigma_{gg \rightarrow (n-2) g} \propto \sigma_{gg \rightarrow gg} \; [\alpha_s \ln^2 (s/s_{cut})]^{n-4} \end{equation} where $s_{cut}$ is the cutoff for the mininum binary invariant $(p_i+p_j)^2 > s_{cut}$ of the 4-momenta of each gluon pair. In the present problem, $s_{cut}=m^2_D$, the double logarithm is not large and certainly does not compensate for the smallness of $\alpha_s$. However, as we have mentioned at the beginning, the collision term on the r.h.s. of \eref{eq:baymeq} consists of the sum of the differences of the reactions in a QCD medium going forward and backward, so a large cross-section does not automatically imply dominance of the corresponding process in the approach to equilibrium. \begin{figure} \centerline{ {\psfig{figure=fig7.ps,width=6.7in}} } \caption{The time development of the same contributions to the total gluon collision entropy density rate as in \fref{gr:s_g-lhc} but at RHIC. The same ratios between the different contributions as at LHC are plotted in (e).} \label{gr:s_g-rhic} \end{figure} \begin{figure} \centerline{ {\psfig{figure=fig8.ps,width=5in}} } \caption{ The time development of the different contributions to the total quark collision entropy density rate at LHC. They are (a) $gg\longleftrightarrow q\bar q$ (b) $gq\longleftrightarrow gq$ or $g\bar q\longleftrightarrow g\bar q$ and (c) the sum of the contributions of all fermion elastic scattering processes $qq\longleftrightarrow qq$, $q\bar q\longleftrightarrow q\bar q$ and $\bar q\bar q\longleftrightarrow \bar q\bar q$. The ratios of the contribution (b) (solid line), (c) (dashed line) to that of (a) is plotted in (d). This shows that throughout the lifetime of the QCD plasma, gluon annihilations into quark-antiquark pairs dominates in the equilibration of the fermions.} \label{gr:s_q-lhc} \end{figure} \noindent Similarly, $gg \longleftrightarrow q\bar q$ is not that different from $gq \longleftrightarrow gq$ or $g\bar q \longleftrightarrow g\bar q$ because the two matrix elements are related simply by a swapping of the Mandelstam variables. So why should the first dominates over the second? Except the different ways that the infrared divergences are cut off in the processes, the main reason is $gg \longrightarrow q\bar q$ dominates over the backward reaction $q\bar q \longrightarrow gg$ due to the simple fact that there are less fermions than gluons present in the plasma. An extreme example of this phenomenon would be the forward and backward reaction balance out each other for all the elastic interactions as in a kinetically equilibrated plasma when only inelastic processes remain in the collision terms. In this extreme, all the ratios of elastic to inelastic collision entropy rate vanish. We can now return to the question of whether other inelastic processes such as $gg \longleftrightarrow q\bar qg$, $gq \longleftrightarrow gqg$, $g\bar q \longleftrightarrow g\bar qg$, $gq \longleftrightarrow qq\bar q$, $g\bar q \longleftrightarrow q\bar q\bar q$, $qq \longleftrightarrow qqg$ etc. should be included. Although they are non-leading compared to $gg \longleftrightarrow ggg$ and $gg \longleftrightarrow q\bar q$ due to colour, they should be significant when one sizes them with the elastic processes in view of the cancellation between the forward and backward reactions. In \cite{wong}, the question of the dominance of inelastic over elastic processes was raised. Here it is sufficient to include the two leading inelastic processes to show this explicitly. Had one included these other processes, then equilibration should be faster and one could end up with a more reasonable quark-antiquark content in the plasma. However, we are doubtful that the equilibration time can be reduced dramatically from what we have shown here. \begin{figure} \centerline{ {\psfig{figure=fig9.ps,width=5in}} } \caption{The time development of the same contributions to the total quark collision entropy density rate as in \fref{gr:s_q-lhc} but at RHIC. The same ratios as at LHC are plotted in (d). They show again inelastic process dominates.} \label{gr:s_q-rhic} \end{figure} As we argued in \cite{wong}, it is hard to perturb a parton system from thermal equilibrium without doing so chemically. Therefore inelastic processes are always active in the approach to equilibrium whereas the same is not true for elastic processes. From our figures, it can be seen that inelastic processes are not there only for chemical equilibration or for minor contributions to thermalization as is commonly assumed due to their possible higher powers in $\alpha_s$, they contribute even more significantly to equilibration than elastic processes. Changing the initial conditions will only vary the dominancy but not remove the dominance. Before closing, we would like to point out some differences of our results with that of PCM. In PCM, there appears to be no early momentaneous isotropic particle momentum distribution in either S+S or Au+Au collisions. The first time that there is approximate isotropy, it is already thermalization according to \cite{geig}. It was claimed that there was no further significant change in the total momentum distribution after $\tau=2.4$ fm/c for Au+Au collision at RHIC. We assume that they mean the shape of the distribution with the exception of the slope which should continue to change due to cooling. However, when the total distribution is broken down into that of the parton components, the approximate isotropy or thermalization becomes less obvious. We have shown that thermalization in the strict sense is slow and isotropy of gluon momentum distribution can be argued to be approximate but that of the fermions is not so good. As to chemical equilibration, PCM shows little chance of that for the fermions. The corresponding fugacity estimates are approaching the ``wrong direction'' with increasing time. This is due to a net outflow of particles from the defined central region. The net flux of outgoing particles is arguably more important for fermions than for gluons because the formers have a larger mean free path. The result is the gluon (fermion) fraction of the particle composition rises (drops) with increasing time. Therefore even if there is no phase transition and the parton plasma is allowed to continue its one-dimensional expansion indefinitely, chemical equilibration will never be achieved. Then according to PCM, the expansion is slow enough for kinetic equilibration for all particle species but too fast for chemical equilibration of the quarks and antiquarks. The boundary effect is too important and is affecting equilibration. In our case, this effect is not incorporated. Although equilibration is slow, full equilibration will be reached given sufficient time. We find it surprising that although the gluon fugacity estimate in PCM \cite{geig&kap} overshoots and stays above or at $1.0$ nearly all the time except at the beginning, $R_g$ is still positive or an order of magnitude larger than $R_q+R_{\bar q}$ when the fugacities of the latter are well below $1.0$ and decreasing. One would expect rather gluon absorption or conversion into quark-antiquark should take a significant toll on the gluon production so that there should be a diminution of gluons. At least, this should be the case when local kinetic equilibrium has been or nearly been reached which PCM claimed to be so at the end of the program run but this is not the case in the plot of the production rate of the different particle species! This is counter-intuitive and opposite to what we have shown. To conclude, we have shown that inelastic processes dominate in the approach towards equilibrium. In particular, gluon branching is most important. Gluon-gluon annihilation into quark-antiquark becomes more important only when the gluons are near saturation and equilibrium. The lower power in $\alpha_s$ of the gluon-gluon elastic scattering as compared to the inelastic gluon emission process is more than compensated for by the cancellation of the reaction going forward and backward. The recovery of isotropy in momentum distribution is slow and so is chemical equilibration. The latter is partly due to the small initial fugacities that we used. As an intrinsic feature of perturbative QCD, the quarks and antiquarks are lagging behind the gluons in equilibration and hence a two-stage equilibration scenario. \section*{Acknowledgements} The author would like to thank M. Fontannaz, D. Schiff and everyone at Orsay for kind hospitality during his stay there, R.D. Pisarski and A.K. Rebhan for raising interesting questions. Thanks also go to R. Baier and everyone at Bielefeld for hospitality during the author's short stay there where this work is completed. The author acknowledges financial support from the Leverhulme Trust.
proofpile-arXiv_065-440
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} The method of effective field theory has repeatedly been used in the analysis of the symmetry breaking sector of the Standard Model. It provides a model independent parametrization of various scenarios for the spontaneous breakdown of the electroweak symmetry. The unknown physics is then hidden in the low energy constants of an effective Lagrangian. The physics in the low energy region of a full theory is adequately described by an effective field theory if corresponding Green functions in both theories have the same low energy structure. One can take this matching requirement as the definition of the effective field theory. It determines functional relationships between the low energy constants of the effective Lagrangian and the parameters of the underlying theory. The special role of gauge theories is readily understood. The effective field theory analysis should not make any particular assumptions about the underlying theory -- apart from symmetry properties and the existence of a mass gap. This also requires parametrizing low energy phenomenology by a gauge-invariant effective Lagrangian. The definition of Green's functions, on the other hand, usually does not reflect the symmetry properties of a gauge theory. Gauge invariance is broken, and the off-shell behaviour of Green's functions is gauge-dependent. Therefore, if the Green functions which enter the matching relations do not reflect the symmetry properties of the full theory, the effective field theory will also include the corresponding gauge artifacts. Without resolving these issues, different approaches to determine the low energy constants at order $p^4$ for the Standard Model with a heavy Higgs boson where presented in several recent articles~\cite{SM_Heavy_Higgs}. For Higgs masses below about~$1 TeV$ the effective Lagrangian can be evaluted explicitly with perturbative methods. These works show clearly that the matching of gauge-dependent quantities causes all kinds of trouble. This calls for a new manifestly gauge-invariant technique which avoids these problems, yet maintains the simplicity and elegance of matching Green's functions as in the ungauged case. \section{A gauge-invariant approach} Any approach to determine the effective Lagrangian for a given underlying gauge theory should match only gauge-invariant quantities. Then one does not have to worry about any gauge artifacts which otherwise might enter the effective field theory. Perhaps the most straightforward idea that comes to mind is to match only $S$-matrix elements. However, this approach is quite cumbersome. In particular, it involves a detailed treatment of infrared physics. Functional techniques like those described in Ref.~\cite{LSM} provide a much easier approach. In this case one matches the generating functionals of Green's functions in the full and the effective theory. Infrared physics drops out at a very early step of the calculation. The remaining contributions all involve the propagation of heavy particles over short distances. Hence, they can be evaluated with a short distance expansion. The computation of loop-integrals is not necessary. Gauge invariance is broken as soon as Green's functions of gauge-depend\-ent operators are considered. Hence, any manifestly gauge-invariant approach must confine itself to analyze Green's functions of gauge-invariant operators, such as the field strength of an Abelian gauge field or the density of the Higgs field. In the following we summarize a manifestly gauge-invariant technique to evaluate the effective Lagrangian describing the low-energy region of the gauged linear sigma model in the spontaneously broken phase. It involves only Green's functions of gauge-invariant operators. For any details the reader is referred to Ref.~\cite{Abelian} where it has been applied to the Abe\-lian case. The Higgs sector of the Standard model with the non-abelian group $SU_L(2)\times U_Y(1)$ can be treated in the same way~\cite{NonAbelian}. We would like to point out, that all $S$-matrix elements of the theory can be evaluated from these Green functions as well. At tree-level the generating functional is given by the classical action. Since the external sources are gauge invariant, i.e., do not couple to the gauge degrees of freedom, the equations of motion can be solved without gauge-fixing. As a result they have a whole class of solutions. Every two representatives are related to each other by a gauge transformation. In order to determine the leading contributions to the low energy constants one merely has to solve the equation of motion for the Higgs boson field. To incorporate higher order corrections one may evaluate the path-integral representation of the generating functional with the method of steepest descent. In this case they are described by Gaussian integrals. Since gauge invariance is manifest, these integrals can also be evaluated without gauge-fixing. As a consequence the gauge degrees of freedom manifest themselves through zero-modes of the quadratic form in the exponent of the gaussian factor. The integration over these modes yields the volume of the gauge group, which can be absorbed by the normalization of the path-integral. The remaining integral over the non-zero modes contains all the physics. We would like to point out one difference between this approach to evaluate a path-integral and the method of Faddeev and Popov: if the gauge is not fixed the evaluation of the path-integral does not involve ghost fields. Hence, the number of diagrams to compute is reduced. The effective Lagrangian of the linear sigma model is a sum of gauge-invariant terms with an increasing number of covariant derivatives and gauge-boson mass factors, corresponding to an expansion in powers of the momentum and the masses. Note that the covariant derivative, the gauge-boson masses and the gauge couplings all count as quantities of order $p$. Thus, the low energy expansion is carried out such that all light-particle singularities are correctly reproduced. Furthermore, if the coupling $\lambda$ of the scalar field is small enough, the low energy constants in the effective Lagrangian admit an expansion in powers of this quantity, corresponding to the loop expansion in the full theory. $n$-loop Feynman diagrams in the Abelian Higgs model yield corrections of order $\lambda^{n-1}$ to the low energy constants. \newpage \section*{Acknowledgments} Work supported in part by Deutsche Forschungsgemeinschaft Grant Nr.\ Ma 1187/7-2, DOE grant \#DE-FG02-92-ER40704, NSF grant PHY-92-18167, and by Schweizerischer Nationalfonds. \section*{References}
proofpile-arXiv_065-441
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{INTRODUCTION} In the last four years the Ape group has been extensively studying lattice QCD in the quenched approximation. Several simulations have been done to study weak matrix elements such as $f_D$, $f_B$, $B_K$ and to study semileptonic decays \cite{latI}-\cite{latV}. These simulations have allowed to study the dependence of the results on the spacing $a$ and to investigate finite size effects. Here we present results for light meson masses and decay constants and baryon spectroscopy. The results have been obtained from eight sets of data at $\beta=6.0$, $6.2$ and $6.4$ using either the Wilson action or the ``improved'' SW-Clover action, in order to reduce $O(a)$ effects. The parameters used in each simulation are listed in Table~\ref{tab:latparams}. The Ape group has performed extensive comparisons on data extracted from smeared and non-smeared propagators and found no real improvement for lattices with a time extent of 64 at $\beta=6.0$ and $\beta=6.2$ \cite{latII_a2}. Note that the simulations have been performed at $\beta$ values of $6.0$ or larger. \setlength{\tabcolsep}{.65pc} \begin{table} \caption{Predicted meson masses in GeV for all lattices.} \label{tab:mesonsGeV} \begin{tabular}{llll} \hline & $M_\rho$ & $M_{\eta'}$ & $M_\phi$ \\ \hline Exper. & 0.770 & 0.686 & 1.019 \\ \hline Lat~I & 0.809(7) & 0.6849(3) & 0.977(7) \\ Lat~II & 0.808(3) & 0.6849(1) & 0.978(3) \\ Lat~III & 0.81(1) & 0.6849(5) & 0.98(1) \\ Lat~IV & 0.803(6) & 0.6851(2) & 0.984(6) \\ Lat~V & 0.79(1) & 0.6856(5) & 1.00(1) \\ Lat~VI & 0.797(7) & 0.6853(3) & 0.989(7)\\ Lat~VII & 0.796(4) & 0.6853(2) & 0.990(4)\\ Lat~VIII& 0.792(4) & 0.6855(2) & 0.994(4)\\ \hline \end{tabular} \end{table} This is to negate the large systematic error present in lattice data for $\beta \;\raisebox{-.5ex}{\rlap{$\sim$}} \raisebox{.5ex}{$<$}\; 6.0$ due to lattice artifacts \cite{chris}. All the results we obtained will be described in greater detail in a forthcoming paper \cite{lavoro}. \section{MESON MASSES AND DECAY CONSTANTS} Meson masses and decay constants have been extracted from two-point correlation functions of the following local operators \[ P_5(x) = i\bar{q}(x)\gamma_5q(x), \;\; V_k(x)=\bar{q}(x)\gamma_k q(x)\; , \] \[ A_\mu(x) = \bar{q}(x)\gamma_\mu\gamma_5q(x) \] in the standard APE way \cite{lavoro}. We fit the correlation functions of these operators to a single particle propagator with a $sinh$ in the case of axial-pseudoscalar function and with a $cosh$ in other cases. \begin{table*}[hbt] \setlength{\tabcolsep}{.55pc} \newlength{\digitwidth} \settowidth{\digitwidth}{\rm 0} \catcode`?=\active \def?{\kern\digitwidth} \label{tab:latparams} \caption{Summary of the parameters of the runs analyzed in this work and time windows used in the fits.} \begin{tabular}{lllllllll} \hline &Lat~I&Lat~II&Lat~III&Lat~IV&Lat~V&Lat~VI&Lat~VII&Lat~VIII\\ \hline Ref & \cite{latI}& \cite{latI,latII_a} & \cite{latIII} & \cite{latIII} & \cite{latV} & \cite{latII_a,latII_a2} & \cite{latIII} & \cite{latIII} \\ $\beta$&$6.0$&$6.0$&$6.2$ &$6.2$&$6.2$&$6.2$&$6.4$&$6.4$\\ Action & SW & Wil & SW & Wil& SW & Wil & Wil & SW \\ \# Confs&200&320&250&250&200&110&400&400\\ Volume&$18^3\times 64$&$18^3\times 64$&$24^3\times 64$& $24^3\times 64$ &$18^3\times 64$&$24^3\times 64$&$24^3\times 64$&$24^3\times 64$\\ \hline $K$& - & - &0.14144&0.1510& - & - &0.1488&0.1400\\ &0.1425&0.1530&0.14184&0.1515& 0.14144&0.1510&0.1492&0.1403\\ &0.1432&0.1540&0.14224&0.1520& 0.14190&0.1520&0.1496&0.1406\\ &0.1440&0.1550&0.14264&0.1526& 0.14244&0.1526&0.1500&0.1409\\ \hline \multicolumn{9}{c}{Mesons with zero momentum} \\ $t_1 - t_2$ & 15-28 & 15-28 & 18-28 & 18-28 & 18-28 & 18-28 & 24-30 & 24-30 \\ \hline \multicolumn{9}{c}{Baryons with zero momentum} \\ $t_1 - t_2$ & 12-21 & 12-21 & 18-28 & 18-28 & 18-28 & 18-28 & 22-28 & 22-28 \\ \hline \end{tabular} \end{table*} The pseudoscalar decay constant $f_{PS}$ has been extracted by combining the fit of $\langle A_0P_5\rangle$ with the ratio $\langle A_0P_5 \rangle /\langle P_5P_5\rangle $. The errors have been estimated by a jacknife procedure, blocking the data in groups of 10 configurations and we have checked that there are no relevant changes in the error estimate by blocking groups of configurations of different size. We have fitted the correlation functions in time windows reported in Table~\ref{tab:latparams}. The time fit intervals are chosen with the following criteria: we fix the lower limit of the intervals as the one at which there is a stabilization of the effective mass and as the upper limit the furthest possible point before the error overwhelms the signal. We discard the possibility of fitting in a restricted region where a plateau is present, as the definition of such a region is highly questionable \cite{fukugita}. For lattices with highest number of configurations, i.e. LatII, LatVII and LatVIII, we confirm that higher statistics do not lead to a longer or better (relative to the statistical error) plateau \cite{fukugita}. \setlength{\tabcolsep}{.3pc} \begin{table} \caption{Extrapolated/interpolated meson decay constants} \label{tab:decaysGeV} \begin{tabular}{llllll} \hline &$\displaystyle\frac{f_\pi}{Z_A m_\rho}$& $\displaystyle\frac{1}{f_\rho Z_V}$&$\displaystyle\frac{f_K}{Z_A m_{K^*}}$ &$\displaystyle\frac{1}{f_{K^*} Z_V}$& \\ \hline Lat~I & 0.17(1) & 0.42(3) & 0.172(9) & 0.39(2) & \\ Lat~II & 0.25(1) & 0.51(2) & 0.239(8) & 0.48(2) & \\ Lat~III & 0.16(1) & 0.39(3) & 0.164(9) & 0.36(2) & \\ Lat~IV & 0.21(1) & 0.47(2) & 0.214(8) & 0.45(1) & \\ Lat~V & 0.19(3) & 0.30(4) & 0.18(2) & 0.30(3) & \\ Lat~VI & 0.21(1) & 0.49(3) & 0.21(1) & 0.46(2) & \\ Lat~VII & 0.23(2) & 0.39(2) & 0.22(1) & 0.38(1) & \\ Lat~VIII& 0.19(1) & 0.30(2) & 0.18(1) & 0.29(1) & \\ \hline \end{tabular} \end{table} Once the hadronic correlation functions have been fitted and the lattice masses and matrix elements extracted, we extract as much physics as possible from the ``strange'' region so that the chiral extrapolation will be needed only in few cases. The method we use is outlined below:\\ {\bf -} We define the lattice planes for meson masses and decay constants ($M_V a$, $(M_{PS} a)^2$), ($f_{PS} a/Z_A$, $(M_{PS} a)^2$) and ($1/(f_{V} Z_V)$, $(M_{PS} a)^2$) where the subscripts $PS$ and $V$ stand for pseudoscalar and vector meson. We plot the Monte Carlo data for each kappa used in the simulation on these planes;\\ {\bf -} On the vector meson plane ($M_V a$, $(M_{PS} a)^2$) we impose the physical ratios $M_{K^*}/M_{K}$, $M_{\rho}/M_{\pi}$ and find the values of $M_\pi a$, $M_\rho a$ (only one independent), $M_K a$, $M_{K^*} a$ (only one independent), $M_{\eta '}a$ and $M_\phi a$;\\ {\bf -} We now use the value of meson masses determined above to read off the lattice meson decay constants, $(f_\pi a/Z_A)$, $(f_K a/Z_A)$, $(f_{\rho} Z_V)^{-1}$ and $(f_{K^*} Z_V)^{-1}$ from the corresponding $f_{PS}$ and $f_V$ planes. \\ This procedure to extract physical quantities only requires meson masses and not unphysical quantities such as quark masses or $k$ values. It allows us to study the $strange$ physics and fix the lattice spacing directly in the region where data have been simulated without chiral extrapolation to zero quark mass. This approach therefore reduces the errors on physical quantities induced by the chiral extrapolation. Using the values of $a^{-1}$ from $M_{K^*}$ we have obtained the physical value in GeV of meson masses reported in Table~\ref{tab:mesonsGeV}. Comparing the vector meson mass (in lattice units) from lattices LatIII, LatV and \cite{ukqcd93} we infer that there is the possibility of some residual finite volume effects on the $18^3$ lattice at $\beta=6.2$ \cite{lavoro}. This problem may also be present in our $\beta=6.4$ data for which the physical volume is the same as in Lat V. Further investigations at larger lattice sizes are necessary to make the situation clearer. Turning to the continuum limit, any dependence of the meson spectrum on $a$ is small and difficult to interpret unambiguously at this stage. In Table~\ref{tab:decaysGeV} we report results for the meson decay constants without including the renormalization constants $Z_V$ and $Z_A$. For both the pseudoscalar and vector decay constants we notice a difference between the Wilson and SW-Clover data. This is presumably due to the different renormalization constants and the smaller $O(a)$ effects in the latter case. There may also be a small residual finite lattice spacing effect in the vector decay constant in the SW-Clover data which needs further study. Overall our results agree with experimental data to $\sim 5\%$ for meson spectrum and to $\sim 10\%-15\%$ for the pseudoscalar decay constants. In our opinion the vector decay constant deserves a much more careful study at larger volume and $\beta$. \section{BARYON MASSES} Baryon masses have been extracted from two-point correlation functions of the following local operators \begin{eqnarray} N & = &\epsilon_{abc}(u^aC\gamma_5 d^b)u^c\nonumber\\ \Delta_\mu & = & \epsilon_{abc}(u^aC\gamma_\mu u^b)u^c\nonumber \end{eqnarray} in the standard way by fitting the two point correlation functions to a single particle propagator with an $exp$ function. The errors have been estimated as in the meson case. \setlength{\tabcolsep}{.19pc} \begin{table} \caption{Predicted baryon masses in GeV for all lattices.} \label{tab:baryonsGeV} \begin{tabular}{llllll} \hline & $M_N $ & $M_{\Lambda\Sigma}$ & $M_\Xi $& $M_\Delta $ & $M_\Omega $\\ \hline Exper. & 0.9389 & 1.135 & 1.3181 & 1.232 & 1.6724 \\ \hline Lat~I & 1.09(5) & 1.21(4) & 1.32(4) & 1.3(1) & 1.60(9) \\ Lat~II & 1.19(5) & 1.29(4) & 1.40(4) & 1.46(7) & 1.71(4) \\ Lat~III & 1.1(1) & 1.22(8) & 1.34(7) & - & - \\ Lat~IV & 1.17(7) & 1.28(6) & 1.39(5) & - & - \\ Lat~V & 1.1(2) & 1.2(2) & 1.4(1) & 1.6(3) & 1.9(2) \\ Lat~VI & 1.2(1) & 1.3(1) & 1.40(9) & 1.50(9) & 1.72(5) \\ Lat~VII & 1.21(9) & 1.32(8) & 1.43(6) & 1.4(2) & 1.72(9) \\ Lat~VIII& 1.2(1) & 1.29(8) & 1.41(7) & 1.3(2) & 1.7(1) \\ \hline \end{tabular} \end{table} We have used the value of meson masses to read off the lattice baryon masses from the planes ($M_N a$, $(M_{PS} a)^2$) and ($M_\Delta a$, $(M_{PS} a)^2$). Using the same values of $a^{-1}$ used for meson masses we have obtained the results reported in Table~\ref{tab:baryonsGeV}. For baryons we find very good agreement with the old APE \cite{parisi} data, while we find slightly larger values when comparing with JLQCD \cite{fukugita} and LANL \cite{gupta}. Also for baryon masses we can conclude that we do not see a dependence on $a$ and that we have an agreement with the experimental data of $\sim 10\%-15\%$.
proofpile-arXiv_065-442
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} Almost all astronomers will agree that most of the mass in the Universe is nonluminous. The nature of this dark matter remains one of the great mysteries of science today. Dynamics of cluster of galaxies suggest a universal nonrelativistic-matter density of $\Omega_0\simeq0.1-0.3$. If the luminous matter were all there was, the duration of the epoch of structure formation would be very short, thereby requiring (in almost all theories of structure formation) fluctuations in the microwave background which would be larger than those observed. These considerations imply $\Omega_0\gtrsim0.3$ \cite{marccmb}. Second, if the current value of $\Omega_0$ is of order unity today, then at the Planck time it must have been $1\pm10^{-60}$ leading us to believe that $\Omega_0$ is precisely unity for aesthetic reasons. A related argument comes from inflationary cosmology, which provides the most satisfying explanation for the smoothness of the microwave background \cite{inflation}. To account for this isotropy, inflation must set $\Omega$ (the {\it total} density, including a cosmological constant) to unity. \begin{figure}[htbp] \centerline{\psfig{file=rotation.ps,width=3.3in}} \bigskip \caption{ Rotation curve for the spiral galaxy NGC6503. The points are the measured circular rotation velocities as a function of distance from the center of the galaxy. The dashed and dotted curves are the contribution to the rotational velocity due to the observed disk and gas, respectively, and the dot-dash curve is the contribution from the dark halo.} \label{rotationfigure} \end{figure} However, the most robust observational evidence for the existence of dark matter involves galactic dynamics. There is simply not enough luminous matter ($\Omega_{\rm lum}\la0.01$) observed in spiral galaxies to account for their observed rotation curves (for example, that for NGC6503 shown in Fig.~\ref{rotationfigure} \cite{broeils}). Newton's laws imply a galactic dark halo of mass $3-10$ times that of the luminous component. With the simple and plausible assumption that the halo of our galaxy is roughly spherical, one can determine that the local dark-matter density is roughly $\rho_0 \simeq 0.3$ GeV~cm$^{-3}$. Furthermore, the velocity distribution of the halo dark matter should be roughly Maxwell-Boltzmann with a velocity dispersion $\simeq270$ km~s$^{-1}$. On the other hand, big-bang nucleosynthesis suggests that the baryon density is $\Omega_b\la0.1$ \cite{bbn}, too small to account for the dark matter in the Universe. Although a neutrino species of mass ${\cal O}(30\, {\rm eV})$ could provide the right dark-matter density, N-body simulations of structure formation in a neutrino-dominated Universe do a poor job of reproducing the observed structure \cite{Nbody}. Furthermore, it is difficult to see (essentially from the Pauli principle) how such a neutrino could make up the dark matter in the halos of galaxies \cite{gunn}. It appears likely then, that some nonbaryonic, nonrelativistic matter is required. The two leading candidates from particle theory are the axion \cite{axion}, which arises in the Peccei-Quinn solution to the strong-$CP$ problem, and a weakly-interacting massive particle (WIMP), which may arise in supersymmetric (or other) extensions of the standard model \cite{jkg}. Here, I review the axion solution to the strong-$CP$ problem, the astrophysical constraints to the axion mass, and prospects for detection of an axion. I then review the WIMP solution to the dark-matter problem and avenues toward detection. Finally, I briefly discuss how measurements of CMB anisotropies may in the future help determine more precisely the amount of exotic dark in the Universe. \section{Axions} Although supersymmetric particles seem to get more attention in the literature lately, we should not forget that the axion also provides a well-motivated and promising alternative dark-matter candidate \cite{axion}. The QCD Lagrangian may be written \begin{equation} {\cal L}_{QCD} = {\cal L}_{\rm pert} + \theta {g^2 \over 32 \pi^2} G \widetilde{G}, \end{equation} where the first term is the perturbative Lagrangian responsible for the numerous phenomenological successes of QCD. However, the second term (where $G$ is the gluon field-strength tensor and $\widetilde{G}$ is its dual), which is a consequence of nonperturbative effects, violates $CP$. However, we know experimentally that $CP$ is not violated in the strong interactions, or if it is, the level of strong-$CP$ violation is tiny. From constraints to the neutron electric-dipole moment, $d_n \la 10^{-25}$ e~cm, it can be inferred that $\theta \la 10^{-10}$. But why is $\theta$ so small? This is the strong-$CP$ problem. The axion arises in the Peccei-Quinn solution to the strong-$CP$ problem \cite{PQ}, which close to twenty years after it was proposed still seems to be the most promising solution. The idea is to introduce a global $U(1)_{PQ}$ symmetry broken at a scale $f_{PQ}$, and $\theta$ becomes a dynamical field which is the Nambu-Goldstone mode of this symmetry. At temperatures below the QCD phase transition, nonperturbative quantum effects break explicitly the symmetry and drive $\theta\rightarrow 0$. The axion is the pseudo-Nambu-Goldstone boson of this near-global symmetry. Its mass is $m_a \simeq\, {\rm eV}\,(10^7\, {\rm GeV}/ f_a)$, and its coupling to ordinary matter is $\propto f_a^{-1}$. {\it A priori}, the Peccei-Quinn solution works equally well for any value of $f_a$ (although one would generically expect it to be less than or of order the Planck scale). However, a variety of astrophysical observations and a few laboratory experiments constrain the axion mass to be $m_a\sim10^{-4}$ eV, to within a few orders of magnitude. Smaller masses would lead to an unacceptably large cosmological abundance. Larger masses are ruled out by a combination of constraints from supernova 1987A, globular clusters, laboratory experiments, and a search for two-photon decays of relic axions \cite{ted}. One conceivable theoretical difficulty with this axion mass comes from generic quantum-gravity arguments \cite{gravity}. For $m_a\sim10^{-4}$ eV, the magnitude of the explicit symmetry breaking is incredibly tiny compared with the PQ scale, so the global symmetry, although broken, must be very close to exact. There are physical arguments involving, for example, the nonconservation of global charge in evaporation of a black hole produced by collapse of an initial state with nonzero global charge, which suggest that global symmetries should be violated to some extent in quantum gravity. When one writes down a reasonable {\it ansatz} for a term in a low-energy effective Lagrangian which might arise from global-symmetry violation at the Planck scale, the coupling of such a term is found to be extraordinarily small (e.g., $\la 10^{-55}$). Of course, we have at this point no predictive theory of quantum gravity, and several mechanisms for forbidding these global-symmetry violating terms have been proposed \cite{solutions}. Therefore, these arguments by no means ``rule out'' the axion solution. In fact, discovery of an axion would provide much needed clues to the nature of Planck-scale physics. Curiously enough, if the axion mass is in the relatively small viable range, the relic density is $\Omega_a\sim1$ and may therefore account for the halo dark matter. Such axions would be produced with zero momentum by a misalignment mechanism in the early Universe and therefore act as cold dark matter. During the process of galaxy formation, these axions would fall into the Galactic potential well and would therefore be present in our halo with a velocity dispersion near 270 km~s$^{-1}$. Although the interaction of axions with ordinary matter is extraordinarily weak, Sikivie proposed a very clever method of detection of Galactic axions \cite{sikivie}. Just as the axion couples to gluons through the anomaly (i.e., the $G\widetilde{G}$ term), there is a very weak coupling of an axion to photons through the anomaly. The axion can therefore decay to two photons, but the lifetime is $\tau_{a\rightarrow \gamma\gamma} \sim 10^{50}\, {\rm s}\, (m_a / 10^{-5}\, {\rm eV})^{-5}$ which is huge compared to the lifetime of the Universe and therefore unobservable. However, the $a\gamma\gamma$ term in the Lagrangian is ${\cal L}_{a\gamma\gamma} \propto a {\vec E} \cdot {\vec B}$ where ${\vec E}$ and ${\vec B}$ are the electric and magnetic field strengths. Therefore, if one immerses a resonant cavity in a strong magnetic field, Galactic axions which pass through the detector may be converted to fundamental excitations of the cavity, and these may be observable \cite{sikivie}. Such an experiment is currently underway and expects to probe the entire acceptable parameter space within the next five years \cite{axionexperiment}. A related experiment, which looks for excitations of Rydberg atoms, may also find dark-matter axions \cite{rydberg}. Although the sensitivity of this technique is supposed to be excellent, it can only cover a limited axion-mass range. It should be kept in mind that there are no accelerator tests for axions in the acceptable mass range. Therefore, these dark-matter axion experiment are actually our {\it only} way to test the Peccei-Quinn solution. \section{Weakly-Interacting Massive Particles} Suppose that in addition to the known particles of the standard model, there exists a new, yet undiscovered, stable (or long-lived) weakly-interacting massive particle (WIMP), $\chi$. At temperatures greater than the mass of the particle, $T\gg m_\chi$, the equilibrium number density of such particles is $n_\chi \propto T^3$, but for lower temperatures, $T\ll m_\chi$, the equilibrium abundance is exponentially suppressed, $n_\chi \propto e^{-m_\chi/T}$. If the expansion of the Universe were so slow that thermal equilibrium was always maintained, the number of WIMPs today would be infinitesimal. However, the Universe is not static, so equilibrium thermodynamics is not the entire story. \begin{figure}[htbp] \centerline{\psfig{file=yyy.ps,width=3.3in}} \bigskip \caption{ Comoving number density of a WIMP in the early Universe. The dashed curves are the actual abundance, and the solid curve is the equilibrium abundance.} \label{YYY} \end{figure} At high temperatures ($T\gg m_\chi$), $\chi$'s are abundant and rapidly converting to lighter particles and {\it vice versa} ($\chi\bar\chi\leftrightarrow l\bar l$, where $l\bar l$ are quark-antiquark and lepton-antilepton pairs, and if $m_\chi$ is greater than the mass of the gauge and/or Higgs bosons, $l\bar l$ could be gauge- and/or Higgs-boson pairs as well). Shortly after $T$ drops below $m_\chi$ the number density of $\chi$'s drops exponentially, and the rate for annihilation of $\chi$'s, $\Gamma=\VEV{\sigma v} n_\chi$---where $\VEV{\sigma v}$ is the thermally averaged total cross section for annihilation of $\chi\bar\chi$ into lighter particles times relative velocity $v$---drops below the expansion rate, $\Gamma\mathrel{\mathpalette\fun <} H$. At this point, the $\chi$'s cease to annihilate, they fall out of equilibrium, and a relic cosmological abundance remains. Fig.~\ref{YYY} shows numerical solutions to the Boltzmann equation which determines the WIMP abundance. The equilibrium (solid line) and actual (dashed lines) abundances per comoving volume are plotted as a function of $x\equiv m_\chi/T$ (which increases with increasing time). As the annihilation cross section is increased the WIMPs stay in equilibrium longer, and we are left with a smaller relic abundance. An approximate solution to the Boltzmann equation yields the following estimate for the current cosmological abundance of the WIMP: \begin{equation} \Omega_\chi h^2={m_\chi n_\chi \over \rho_c}\simeq \left({3\times 10^{-27}\,{\rm cm}^3 \, {\rm sec}^{-1} \over \sigma_A v} \right), \label{eq:abundance} \end{equation} where $h$ is the Hubble constant in units of 100 km~s$^{-1}$~Mpc$^{-1}$. The result is to a first approximation independent of the WIMP mass and is fixed primarily by its annihilation cross section. The WIMP velocities at freeze out are typically some appreciable fraction of the speed of light. Therefore, from equation~(\ref{eq:abundance}), the WIMP will have a cosmological abundance of order unity today if the annihilation cross section is roughly $10^{-9}$ GeV$^{-2}$. Curiously, this is the order of magnitude one would expect from a typical electroweak cross section, \begin{equation} \sigma_{\rm weak} \simeq {\alpha^2 \over m_{\rm weak}^2}, \end{equation} where $\alpha \simeq {\cal O}(0.01)$ and $m_{\rm weak} \simeq {\cal O}(100\, {\rm GeV})$. The value of the cross section in equation~(\ref{eq:abundance}) needed to provide $\Omega_\chi\sim1$ comes essentially from the age of the Universe. However, there is no {\it a priori} reason why this cross section should be of the same order of magnitude as the cross section one would expect for new particles with masses and interactions characteristic of the electroweak scale. In other words, why should the age of the Universe have anything to do with electroweak physics? This ``coincidence'' suggests that if a new, yet undiscovered, massive particle with electroweak interactions exists, then it should have a relic density of order unity and therefore provides a natural dark-matter candidate. This argument has been the driving force behind a vast effort to detect WIMPs in the halo. The first WIMPs considered were massive Dirac or Majorana neutrinos with masses in the range of a few GeV to a few TeV. (Due to the Yukawa coupling which gives a neutrino its mass, the neutrino interactions become strong above a few TeV, and it no longer remains a suitable WIMP candidate \cite{unitarity}.) LEP ruled out neutrino masses below half the $Z^0$ mass. Furthermore, heavier Dirac neutrinos have been ruled out as the primary component of the Galactic halo by direct-detection experiments (described below) \cite{heidelberg}, and heavier Majorana neutrinos have been ruled out by indirect-detection experiments \cite{kamiokande} (also described below) over much of their mass range. Therefore, Dirac neutrinos cannot comprise the halo dark matter \cite{griestsilk}; Majorana neutrinos can, but only over a small range of fairly large masses. This was a major triumph for experimental particle astrophysicists:\ the first falsification of a dark-matter candidate. However, theorists were not too disappointed: The stability of a fourth generation neutrino had to be postulated {\it ad hoc}---it was not guaranteed by some new symmetry. So although heavy neutrinos were plausible, they certainly were not very well-motivated from the perspective of particle theory. A much more promising WIMP candidate comes from supersymmetry (SUSY) \cite{jkg,haberkane}. SUSY was hypothesized in particle physics to cure the naturalness problem with fundamental Higgs bosons at the electroweak scale. Coupling-constant unification at the GUT scale seems to be improved with SUSY, and it seems to be an essential ingredient in theories which unify gravity with the other three fundamental forces. As another consequence, the existence of a new symmetry, $R$-parity, in SUSY theories guarantees that the lightest supersymmetric particle (LSP) is stable. In the minimal supersymmetric extension of the standard model (MSSM), the LSP is usually the neutralino, a linear combination of the supersymmetric partners of the photon, $Z^0$, and Higgs bosons. (Another possibility is the sneutrino, but these particles interact like neutrinos and have been ruled out over most of the available mass range \cite{sneutrino}.) Given a SUSY model, the cross section for neutralino annihilation to lighter particles is straightforward, so one can obtain the cosmological mass density. The mass scale of supersymmetry must be of order the weak scale to cure the naturalness problem, and the neutralino will have only electroweak interactions. Therefore, it is to be expected that the cosmological neutralino abundance is of order unity. In fact, with detailed calculations, one finds that the neutralino abundance in a very broad class of supersymmetric extensions of the standard model is near unity and can therefore account for the dark matter in our halo \cite{ellishag}. If neutralinos reside in the halo, there are several avenues for detection \cite{jkg}. One of the most promising techniques currently being pursued involves searches for the ${\cal O}({\rm keV})$ recoils produced by elastic scattering of neutralinos from nuclei in low-background detectors \cite{witten,labdetectors}. Another strategy is observation of energetic neutrinos produced by annihilation of neutralinos in the Sun and Earth in converted proton-decay and astrophysical-neutrino detectors (such as MACRO, Kamiokande, IMB, AMANDA, and NESTOR) \cite{SOS}. There are also searches for anomalous cosmic rays which would be produced by annihilation of WIMPs in the halo. Of course, SUSY particles should also show up in accelerator searches if their mass falls within the experimentally accessible range. Although supersymmetry provides perhaps the most promising dark-matter candidate (and solves numerous problems in particle physics), a practical difficulty with supersymmetry is that we have little detailed predictive power. In SUSY models, the standard-model particle spectrum is more than doubled, and we really have no idea what the masses of all these superpartners should be. There are also couplings, mixing angles, etc. Therefore, what theorists generally do is survey a huge set of models with masses and couplings within a plausible range, and present results for relic abundances and direct- and indirect-detection rates, usually as scatter plots versus neutralino mass. Energetic neutrinos from WIMP annihilation in the Sun or Earth would be inferred by observation of neutrino-induced upward muons coming from the direction of the Sun or the core of the Earth. Predictions for the fluxes of such muons in SUSY models seem to fall for the most part between $10^{-6}$ and 1 event~m$^{-2}$~s$^{-1}$ \cite{jkg}, although the numbers may be a bit higher or lower in some models. Presently, IMB and Kamiokande constrain the flux of energetic neutrinos from the Sun to be less than about 0.02 m$^{-2}$~s$^{-1}$ \cite{kamiokande,imb}. MACRO expects to be able to improve on this sensitivity by perhaps an order of magnitude. Future detectors may be able to improve on this limit further. For example, AMANDA expects to have an area of roughly $10^4$ m$^2$, and a $10^6$-m$^2$ detector is being discussed. However, it should be kept in mind that without muon energy resolution, the sensitivity of these detectors will not approach the inverse exposure; it will be limited by the atmospheric-neutrino background. If a detector has good angular resolution, the signal-to-noise ratio can be improved, and even moreso with energy resolution, so sensitivities approaching the inverse exposure could be achieved \cite{joakim}. Furthermore, ideas for neutrino detectors with energy resolution are being discussed \cite{wonyong}, although at this point these appear likely to be in the somewhat-distant future. The other possibility is direct detection of a WIMP via observation of the nuclear recoil induced by WIMP-nucleus elastic scattering in a low-background detector. The predicted rates depend on the target nucleus adopted. For example, in a broad range of SUSY models, the predicted scattering rates in a germanium detector seem to fall for the most part between $10^{-4}$ to 10 events~kg$^{-1}$~day$^{-1}$ \cite{jkg}, although again, there may be models with higher or lower rates. Current experimental sensitivities in germanium detectors are around 10 events~kg$^{-1}$~day$^{-1}$ \cite{heidelberg}. To illustrate future prospects, consider the CDMS experiment \cite{cdms} which expects to soon have a kg germanium detector with an background rate of 1 event~day$^{-1}$. After a one-year exposure, their sensitivity would therefore be ${\cal O}(0.1\, {\rm event~kg}^{-1}\,{\rm day}^{-1})$; this could be improved with better background rejection. Future detectors will achieve better sensitivities, and it should be kept in mind that numerous other target nuclei are being considered by other groups. However, it also seems clear that it will be quite a while until a good fraction of the available SUSY parameter space is probed. Generally, most theorists have just plugged in SUSY parameters into the machinery which produces detection rates and plotted results for direct and indirect detection. However, another approach is to compare, in a somewhat model-independent although approximate fashion, the rates for direct and indirect detection \cite{jkg,taorich,bernard}. The underlying observation is that the rates for the two types of detection are both controlled primarily by the WIMP-nucleon coupling. One must then note that WIMPs generally undergo one of two types of interaction with the nucleon: an axial-vector interaction in which the WIMP couples to the nuclear spin (which, for nuclei with nonzero angular momentum is roughly 1/2 and {\it not} the total angular momentum), and a scalar interaction in which the WIMP couples to the total mass of the nucleus. The direct-detection rate depends on the WIMP-nucleon interaction strength and on the WIMP mass. On the other hand, indirect-detection rates will have an additional dependence on the energy spectrum of neutrinos from WIMP annihilation. By surveying the various possible neutrino energy spectra, one finds that for a given neutralino mass and annihilation rate in the Sun, the largest upward-muon flux is roughly three times as large as the smallest \cite{bernard}. So even if we assume the neutralino-nucleus interaction is purely scalar or purely axial-vector, there will still be a residual model-dependence of a factor of three when comparing direct- and indirect-detection rates. For example, for scalar-coupled WIMPs, the event rate in a kg germanium detector will be equivalent to the event rate in a $(2-6)\times 10^6$ m$^2$ neutrino detector for 10-GeV WIMPs and $(3-5)\times10^4$ m$^2$ for TeV WIMPs \cite{bernard}. Therefore, the relative sensitivity of indirect detection when compared with the direct-detection sensitivity increases with mass. The bottom line of such an analysis seems to be that direct-detection experiments will be more sensitive to neutralinos with scalar interactions with nuclei, although very-large neutrino telescopes may achieve comparable sensitivities at larger WIMP masses. This should come as no surprise given the fact that direct-detection experiments rule out Dirac neutrinos \cite{heidelberg}, which have scalar-like interactions, far more effectively than do indirect-detection experiments \cite{bernard}. Generically, the sensitivity of indirect searches (relative to direct searches) should be better for WIMPs with axial-vector interactions, since the Sun is composed primarily of nuclei with spin (i.e., protons). However, a comparison of direct- and indirect-detection rates is a bit more difficult for axially-coupled WIMPs, since the nuclear-physics uncertainties in the neutralino-nuclear cross section are much greater, and the spin distribution of each target nucleus must be modeled. Still, in a careful analysis, Rich and Tao found that in 1994, the existing sensitivity of energetic-neutrino searches to axially-coupled WIMPs greatly exceeded the sensitivities of direct-detection experiments \cite{taorich}. To see how the situation may change with future detectors, let us consider a specific axially-coupled dark-matter candidate, the light Higgsino recently put forward by Kane and Wells \cite{kanewells}. In order to explain the anomalous CDF $ee\gamma \gamma + \slashchar{E}_T$ \cite{CDF}, the $Z\rightarrow b\bar b$ anomaly, and the dark matter, this Higgsino must have a mass between 30--40 GeV. Furthermore, the coupling of this Higgsino to quarks and leptons is due primarily to $Z^0$ exchange with a coupling proportional to $\cos 2\beta$, where $\tan\beta$ is the usual ratio of Higgs vacuum expectation values in supersymmetric models. Therefore, the usually messy cross sections one deals with in a general MSSM simplify for this candidate, and the cross sections needed for the cosmology of this Higgsino depend only on the two parameters $m_\chi$ and $\cos2\beta$. Furthermore, since the neutralino-quark interaction is due only to $Z^0$ exchange, this Higgsino will have only axial-vector interactions with nuclei. The Earth is composed primarily of spinless nuclei, so WIMPs with axial-vector interactions will not be captured in the Earth, and we expect no neutrinos from WIMP annihilation therein. However, most of the mass in the Sun is composed of nuclei with spin (i.e., protons). The flux of upward muons induced by neutrinos from annihilation of these light Higgsinos would be $\Gamma_{\rm det}\simeq 2.7\times10^{-2}\, {\rm m}^{-2}\, {\rm yr}^{-1}\, \cos^2 2\beta$ \cite{katie}. On the other hand, the rate for scattering from $^{73}$Ge is $R\simeq 300\, \cos^2 2\beta\, {\rm kg}^{-1}\, {\rm yr}^{-1}$ \cite{kanewells,katie}. For illustration, in addition to their kg of natural germanium, the CDMS experiment also plans to run with 0.5 kg of (almost) purified $^{73}$Ge. With a background event rate of roughly one event~kg$^{-1}$~day$^{-1}$, after one year, the $3\sigma$ sensitivity of the experiment will be roughly 80 kg$^{-1}$~yr$^{-1}$. Comparing the predictions for direct and indirect detection of this axially-coupled WIMP, we see that the enriched-$^{73}$Ge sensitivity should improve on the {\it current} limit to the upward-muon flux ($0.02$ m$^{-2}$ yr$^{-1}$) roughly by a factor of 4. When we compare this with the forecasted factor-of-ten improvement expected in MACRO, it appears that the sensitivity of indirect-detection experiments looks more promising. Before drawing any conclusions, however, it should be noted that the sensitivity in detectors with other nuclei with spin may be significantly better. On the other hand, the sensitivity of neutrino searches increases relative to direct-detection experiments for larger WIMP masses. It therefore seems at this point that the two schemes will be competitive for detection of light axially-coupled WIMPs, but the neutrino telescopes may have an advantage in probing larger masses. A common question is whether theoretical considerations favor a WIMP which has predominantly scalar or axial-vector couplings. Unfortunately, there is no simple answer. When detection of supersymmetric dark matter was initially considered, it seemed that the neutralino in most models would have predominantly axial-vector interactions. It was then noted that in some fraction of models where the neutralino was a mixture of Higgsino and gaugino, there could be some significant scalar coupling as well \cite{kim}. As it became evident that the top quark had to be quite heavy, it was realized that nondegenerate squark masses would give rise to scalar couplings in most models \cite{drees}. However, there are still large regions of supersymmetric parameter space where the neutralino has primarily axial-vector interactions, and in fact, the Kane-Wells Higgsino candidate has primarily axial-vector interactions. The bottom line is that theory cannot currently reliable say which type of interaction the WIMP is likely to have, so experiments should continue to try to target both. \section{Dark Matter and the Cosmic Microwave Background} The key argument for nonbaryonic dark matter relies on the evidence that the total nonrelativistic-matter density $\Omega_0 \mathrel{\mathpalette\fun >} 0.1$, outweighs the baryon density $\Omega_b\mathrel{\mathpalette\fun <} 0.1$ allowed by big-bang nucleosynthesis. With the advent of a new generation of long-duration balloon-borne and ground-based interferometry experiments and NASA's MAP \cite{MAP} and ESA's COBRAS/SAMBA \cite{COBRAS} missions, CMB measurements will usher in a new era in cosmology. In forthcoming years, the cosmic microwave background (CMB) may provide a precise inventory of the matter content in the Universe and confirm the discrepancy between the baryon density and the total nonrelativistic-matter density, if it indeed exists. The primary goal of these experiments is recovery of the temperature autocorrelation function or angular power spectrum of the CMB. The fractional temperature perturbation $\Delta T({\bf \hat n})/T$ in a given direction ${\bf \hat n}$ can be expanded in terms of spherical harmonics, \begin{equation} {\Delta T({\bf \hat n}) \over T} = \sum_{lm} \, a_{(lm)}\, Y_{(lm)}({\bf \hat n}), \label{Texpansion} \end{equation} where the multipole coefficients are given by \begin{equation} a_{(lm)} = \int\, d{\bf \hat n}\, Y_{(lm)}^*({\bf \hat n}) \, {\Delta T({\bf \hat n}) \over T}. \label{alms} \end{equation} Cosmological theories predict that these multipole coefficients are statistically independent and are distributed with variance $ \VEV{a_{(lm)}^* a_{(l'm')} } = C_l \, \delta_{ll'} \, \delta_{mm'}$. Roughly speaking, each $C_l$ measures the square of the mean temperature difference between two points separated by an angle $\theta\sim \pi/l$. \begin{figure}[htbp] \centerline{\psfig{file=models.eps,width=3.7in}} \bigskip \caption{ Theoretical predictions for CMB spectra as a function of multipole moment $\ell$ for models with primordial adiabatic perturbations. In each case, the heavy curve is that for the standard-CDM values, a total density $\Omega=1$, cosmological constant $\Lambda=0$, baryon density $\Omega_b=0.06$, and Hubble parameter $h=0.5$. Each graph shows the effect of variation of one of these parameters. In (d), $\Omega=\Omega_0+\Lambda=1$.} \label{curves} \end{figure} Theoretical predictions for the $C_l$'s can be made given a theory for structure formation and the values of several cosmological parameters. For example, Fig.~\ref{curves} shows predictions for multipole moments in models with primordial adiabatic perturbations. The peaks in the spectra come from oscillations in the photon-baryon fluid at the surface of last scatter, and the damping at small angles is due to the finite thickness of the surface of last scatter. Each panel shows the effect of independent variation of one of the cosmological parameters. As illustrated, the height, width, and spacing of the acoustic peaks in the angular spectrum depend on these (and other) cosmological parameters. The CMB spectrum also depends on the model (e.g., inflation or topological defects) for structure formation, the ionization history of the Universe, and the presence of gravity waves. However, no two of the classical cosmological parameters affects the CMB spectrum in precisely the same way. For example, the angular position of the first peak depends primarily on the geometry ($\Omega=\Omega_0 +\Lambda$ where $\Lambda$ is the contribution of the cosmological constant) of the Universe \cite{kamspergelsug}, but is relatively insensitive to variations in the other parameters. Assuming that the primordial perturbations were adiabatic, we could fit for all of these parameters if the angular spectrum could be measured precisely. COBE normalizes the amplitude and slope of the CMB spectrum to $\sim10\%$. However, the angular resolution was not fine enough to probe the detailed shape of the acoustic peaks in the power spectrum, so COBE was unable to capitalize on this wealth of information. Nor can it discriminate between scalar and tensor modes. A collection of recent ground-based and balloon-borne experiments seem to confirm a first acoustic peak, but they still cannot determine its precise height, width, or location. In the next few years, long-duration balloon flights (e.g., BOOMERANG and TOPHAT) and ground-based interferometry experiments (e.g., CAT and CBI) will begin to discern the first and higher few peaks. Subsequently, future satellite experiments, such as NASA's MAP mission \cite{MAP} and then ESA's COBRAS/SAMBA \cite{COBRAS} will accurately map the CMB temperature over most of the sky with good angular resolution and will therefore be able to recover the CMB power spectrum with precision. Of course, the precision attainable is ultimately limited by cosmic variance and practically by a finite angular resolution, instrumental noise, and partial sky coverage in a realistic CMB mapping experiment. Assuming that primordial perturbations are adiabatic, one finds that with future satellite missions, $\Omega$ may potentially be determined to better than 10\% {\it after} marginalizing over all other undetermined parameters, and better than 1\% if the other parameters can be fixed by independent observations or assumption \cite{jkksone}. This would be far more accurate than any traditional determinations of the geometry. (Of course, if primordial perturbations turn out to be isocurvature or due to topological defects, this may not be the case.) The cosmological constant $\Lambda$ will be determined with a similar accuracy, so the nonrelativistic-matter density $\Omega_0$ will also be accurately determined \cite{jkkstwo}. Small variations in the baryon density have a dramatic effect on the CMB spectrum, so $\Omega_b$ will be determined with even greater precision. Therefore, if there is more nonrelativistic matter in the Universe than the baryons can account for, as current evidence suggests, it should become clear with these future CMB experiments. The CMB will also measure the Hubble constant and perhaps be sensitive to a small neutrino mass \cite{gds}. Temperature maps will also begin to disentangle the scalar and tensor (i.e., long-wavelength gravity-wave) contributions to the CMB and determine their primordial spectra, and this could be used to test inflation \cite{jkkstwo}. CMB polarization maps may also help isolate the tensor contribution \cite{polarization}. Therefore, the CMB will become an increasingly powerful probe of the early Universe. \acknowledgements{This work was supported by the D.O.E. under contract DEFG02-92-ER 40699, NASA under NAG5-3091, and the Alfred P. Sloan Foundation.}
proofpile-arXiv_065-443
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{INTRODUCTION} A precise theoretical framework is needed for the study of the quark mass effects in physical observables because quarks are not free particles. In fact, the quark masses should be seen more like coupling constants than like physical parameters. The perturbative pole mass and the running mass are the two most commonly used quark mass definitions. The perturbative pole mass, $M(p^2=M^2)$, is defined as the pole of the renormalized quark propagator in a strictly perturbative sense. It is gauge invariant and scheme independent. However, it appears to be ambiguous due to non-perturbative renormalons. The running mass, $\bar{m}(\mu)$, the renormalized mass in the $\overline{MS}$ scheme, does not suffer from this ambiguity. Both quark mass definitions can be related perturbatively through \begin{equation} M = \bar{m}(\mu) \left\{ 1 + \frac{\as(\mu)}{\pi} \left[ \frac{4}{3}-\log \frac{\bar{m}^2(\mu)}{\mu^2} \right] \right\}. \label{relates} \end{equation} Heavy quark masses, like the bottom quark mass, can be extracted using QCD Sum Rules or lattice calculations from the quarkonia spectrum, see \cite{UIMP} and references therein. The bottom quark perturbative pole mass appears to be around $M_b = 4.6-4.7 (GeV)$ whereas the running mass at the running mass scale reads $\bar{m}_b(\bar{m}_b) = (4.33 \pm 0.06) GeV$. Performing the running until the $Z$-boson mass scale we find $\bar{m}_b(M_Z) = (3.00 \pm 0.12) GeV$. Since for the bottom quark the difference between the perturbative pole mass and the running mass at the $M_Z$ scale is quite significant it is crucial to specify in any theoretical perturbative prediction at $M_Z$ which mass should we use. The relative uncertainty in the strong coupling constant decreases in the running from low to high energies as the ratio of the strong coupling constants at both scales. On the contrary, if we perform the quark mass running with the extreme mass and strong coupling constant values and take the maximum difference as the propagated error, induced by the strong coupling constant error, the quark mass uncertainty increases following \begin{equation} \varepsilon_r (\bar{m}(M_Z)) \simeq \varepsilon_r (\bar{m}(\mu)) \end{equation} \[ + \frac{2 \gamma_0}{\beta_0} \left(\frac{\as(\mu)}{\as(M_Z)}-1 \right) \varepsilon_r (\as(M_Z)), \] where $\gamma_0=2$ and $\beta_0=11-2/3 N_F$, see figure~\ref{running}. We use the world average \cite{bethke} value $\as = 0.118 \pm 0.006$ for the strong coupling constant. It is interesting to stress, looking at figure~\ref{running}, that even a big uncertainty in a possible evaluation of the bottom quark mass at the $M_Z$ scale can be competitive with low energy QCD Sum Rules and lattice calculations with smaller errors~\footnote{A recent lattice evaluation \cite{lattice} has enlarged the initial estimated error on the bottom quark mass, $\bar{m}_b(\bar{m}_b) = (4.15 \pm 0.20) GeV$, due to unknown higher orders in the perturbative matching of the HQET to the full theory.}. Furthermore, non-perturbative contributions are expected to be negligible at the $Z$-boson mass scale. The running mass holds another remarkable feature. Total cross sections can exhibit potentially dangerous terms of the type $M^2 \log M^2/s$ that however can be absorbed \cite{chety} using Eq.~(\ref{relates}) and expressing the total result in terms of the running mass. \mafigura{5.5cm}{running.ps} {Running of the bottom quark mass from low energies to the $M_Z$ scale. Upper line is the run of $\bar{m}_b(\bar{m}_b)=4.39(GeV)$ with $\as (M_Z)=0.112$. Bottom line is the run of $\bar{m}_b(\bar{m}_b)=4.27(GeV)$ with $\as (M_Z)=0.124$. Second picture is the difference of both, our estimate for the propagated error.} {running} \section{THREE JETS OBSERVABLES AT LO} Quark masses can be neglected for many observables at LEP because usually they appear as the ratio $m_q^2/M_Z^2$. For the heaviest quark produced at LEP, the bottom quark, this means a correction of 3 per mil for a quark mass of 5 (GeV). Even if the coefficient in front is 10 we get at most a 3\% effect, 1\% if we use the bottom quark running mass at $M_Z$. This argument is true for total cross section. However, jet cross sections depend on a new variable, $y_c$, the jet-resolution parameter that defines the jet multiplicity. This new variable introduces a new scale in the analysis, $E_c = M_Z \sqrt{y_c}$, that for small values of $y_c$ could enhance the effect of the quark mass as $m_b^2/E_c^2 = (m_b^2/M_Z^2)/y_c$. The high precision achieved at LEP makes these effects relevant. In particular, it has been shown \cite{Juano} that the biggest systematic error in the measurement of $\as(M_Z)$ from $b\bar{b}$-production at LEP from the ratio of three to two jets comes from the uncertainties in the estimate of the quark mass effects. We are going to study the effect of the bottom quark mass in the following ratios of three-jet decay rates and angular distributions \begin{eqnarray} R_3^{bd} &\equiv& \frac{\Gamma^b_{3j}(y_c)/\Gamma^b} {\Gamma^d_{3j}(y_c)/\Gamma^d}, \label{r3bd}\\ R^{bd}_\vartheta &\equiv& \left. \frac{1}{\Gamma^b}\frac{d\Gamma^b_{3j}}{d\vartheta}\right/ \frac{1}{\Gamma^d}\frac{d\Gamma^d_{3j}}{d\vartheta}, \label{rtheta} \end{eqnarray} where we consider massless the $d$-quark and $\vartheta$ is the minimum of the angles formed between the gluon jet and the quark and antiquark jets. Both observables are normalized to the total decay rates in order to cancel large weak corrections dependent on the top quark mass \cite{top}. \def1.5{1.5} \begin{table}[t] \caption{The jet-clustering algorithms} \label{algorithms} \begin{center} \begin{tabular}{ll} \hline Algorithm & Resolution \\ \hline EM & $2(p_i \cdot p_j)/s$ \\ JADE & $2(E_i E_j)(1-\cos \vartheta_{ij})/s$ \\ E & $(p_i+p_j)^2/s$ \\ DURHAM & $2 \min(E_i^2,E_j^2)(1-\cos \vartheta_{ij})/s$ \\ \hline \end{tabular} \end{center} \end{table} \mafigura{6cm}{LO.ps} {Feynman diagrams contributing to the three-jets decay rate of $Z\rightarrow b\bar{b}$ at order $\as$.} {feynmanLO} At LO in the strong coupling constant we must compute the amplitudes of the Feynman diagrams depicted in figure \ref{feynmanLO} plus the interchange of a virtual gluon between the quark and antiquark that only contributes to the two-jet decay rate. In addition to renormalized UV divergences, IR singularities, either collinear or soft, appear because of the presence of massless particles like gluons. Bloch-Nordsiek and Kinoshita-Lee-Nauemberg theorems\cite{IR} assure IR divergences cancel for inclusive cross section. Technically this means, if we use DR to regularize the IR divergences of the loop diagrams we should express the phase space for the tree-level diagrams in arbitrary $D$-dimensions. The IR singularities cancel when we integrate over the full phase space. Another delicate question is the problem of hadronization. Perturbative QCD gives results at the level of partons, quarks and gluons, but in nature one observes hadrons, not partons, and hadronization can shift the QCD predictions. We apply to the parton amplitudes the same jet clustering algorithms applied experimentally to the real observed particles, see table~\ref{algorithms}. Starting from a bunch of particles of momenta $p_i$ we calculate, for instance, $y_{ij}=2(p_i \cdot p_j)/s$, the scalar product of all the possible momenta pairs. If the minimum is smaller than a fixed $y_c$ we combine the two involved particles in a new pseudoparticle of momentum $p_i+p_j$. The procedure starts again until all the $y_{ij}$ are bigger than $y_c$. The number of pseudoparticles at the end of the procedure defines the number of jets. The jet clustering algorithms automatically define IR finite quantities. For the moment, we do not enter in the question of which is the best jet clustering algorithm although the main criteria followed to choose one of them should be based in two requirements: minimization of higher order corrections and insensitivity to hadronization. If we restrict to the three-jet decay rate the IR problem can be overcome and everything can be calculated in four dimensions because the jet clustering algorithms automatically exclude the IR region from the three-body phase space. For massless particles and at the lowest order the EM~\cite{LO}, JADE and E algorithms give the same answers. Analytical results for the massless three-jet fraction exist for both JADE-like~\cite{KN} and DURHAM~\cite{Durham} algorithms. A complete analysis for the ratios of three-jet decay rates and the angular distributions quoted in Eq.~\ref{r3bd} and \ref{rtheta} can be found in \cite{LO}. For practical purposes a parametrization of the result in terms of a power series in $\log y_c$ gives a good description \cite{KN,LO}. \mafigura{6cm}{NLO.ps} {Feynman diagrams contributing to the three-jets decay rate of $Z\rightarrow b\bar{b}$ at order $\as^2$. Self-energies in external legs have not been shown.} {feynmanNLO} \section{THREE JETS OBSERVABLES AT NLO} The effect of the bottom quark mass has been studied experimentally by~\cite{Joan} on the $R_3^{bd}$ ratio. As we have seen the running of the bottom quark mass from low energies to the $M_Z$ scale is quite strong. The LO QCD prediction for $R_3^{bd}$ does not allow us to distinguish which mass we should use in the theoretical expressions, either the pole mass or the running mass at some scale. The computation of the NLO is mandatory if we want to extract information about the bottom quark mass from LEP data. At the NLO we have to calculate the interference of the loop diagrams depicted in figure~\ref{feynmanNLO} with the lowest order Feynman diagrams of figure~\ref{feynmanLO} plus the square of the tree-level diagrams of figure~\ref{feynmanNLO}. The amplitudes in the massless case were calculated by~\cite{ERT,KL}. The implementation of the jet clustering algorithms was performed by~\cite{KN}. The main problem that now we can not avoid is the appearance of IR singularities. With massive quarks we loose all the quark-gluon collinear divergences. The amplitudes behave better in the IR region. The disadvantage however is the mass itself. We have to perform quite more complicated loop and phase space integrals. Furthermore, we still conserve the gluon-gluon collinear divergences leading to IR double poles. The three-jet decay rate can be written as \begin{equation} \Gamma^{b}_{3j} = C [g_V^2 H_V(y_c,r_b) + g_A^2 H_A(y_c,r_b)], \end{equation} where $r_b=m_b^2/M_Z^2$, $C=M_Z \: g^2/(c_W^2 64 \pi) (\as/\pi)$ is a normalization constant that disappear in the ratio and $g_V$ and $g_A$ are the vector and the axial-vector neutral current quark couplings. At tree-level and for the bottom quark $g_V = -1 + 4 s_W^2/3$ and $g_A = 1$. Now we can expand the functions $H_{V(A)}$ in $\as$ and factorize the leading dependence on the quark mass as follows \begin{eqnarray} H_{V(A)} &=& A^{(0)}(y_c) + r_b B_{V(A)}^{(0)}(y_c,r_b) \\ &+& \frac{\as}{\pi} \left( A^{(1)}(y_c) + r_b B_{V(A)}^{(1)}(y_c,r_b) \right), \nonumber \end{eqnarray} where we have taken into account that for massless quarks vector and axial contributions are identical\footnote{We do not consider the small $O(\as^2)$ triangle anomaly \cite{triangle}. With our choice of the normalization $A^{(0)}(y_c)=A(y_c)/2$ and $A^{(1)}(y_c)=B(y_c)/4$, where $A(y_c)$ and $B(y_c)$ are defined in \cite{KN}.}. First steep in the calculation is to show the cancellation of the IR divergences in order to build matrix elements free of singularities. It is possible to do it analytically. However, we knew from the beginning IR divergences should disappear \cite{IR}. The challenge is in the calculation of the finite parts. This calculation is rather long, complex and full of difficulties. Strong cancellations occur between different groups of diagrams making difficult even a numerical approach. We have taken as guide line the massless result of \cite{ERT,KN} although the IR structure of the massive case is completely different from the massless one. \mafigura{7cm}{Jtest.ps} {NLO vector contribution to the three-jet decay rate of $Z\rightarrow b\bar{b}$ for bottom quark masses from $1$ to $5(GeV)$ and fixed $y_c$ in the JADE algorithm. Big circle is the massless case.} {Jtest} \mafigura{7cm}{Etest.ps} {NLO vector contribution to the three-jet decay rate of $Z\rightarrow b\bar{b}$ for bottom quark masses from $1$ to $5(GeV)$ and fixed $y_c$ in the E algorithm. Big circle is the massless case.} {Etest} In figures \ref{Jtest} and \ref{Etest} we present our preliminary result for the vectorial contribution to the $O(\as^2)$ three-jet decay rate of the Z-boson into bottom quarks. We have performed the calculation for different values of the bottom quark mass from $1$ to $5(GeV)$ for fixed $y_c$. We want to show we can recover the massless result \cite{KN}, depicted as a big circle, i.e., in the limit of massless quarks we reach the $A^{(1)}(y_c)$ function. This is our main test to have confidence in our calculation. In the JADE algorithm we can see that for big values of $y_c$ the NLO corrections due to the quark mass are very small and below the massless result. Notice they increase quite a lot for small values of $y_c$ and give a positive correction that will produce a change in the slope of the LO prediction for $R_3^{bd}$. In any case we recover the massless limit and a linear parametrization in the quark mass squared could provide a good description. The E algorithm behaves also linearly in the quark mass squared although only for big values of $y_c$. Corrections in the E algorithm are always very strong. The reason is the following, the resolution parameter for the E algorithm explicitly incorporates the quark mass, $y_{ij}=(p_i+p_j)/s$, i.e., for the same value of $y_c$ we are closer to the two-jet IR region and the difference from the other algorithms is precisely the quark mass. This phenomenon already manifest at the LO. The behaviour of the E algorithm is completely different from the others for massive quarks. It is difficult to believe in the E algorithm as a good prescription for physical applications since mass corrections as so big. However for the same reason, it seems to be the best one for testing massive calculations. \section{CONCLUSIONS} We have presented the first results for the NLO strong corrections to the three-jet decay rate of the Z-boson into massive quarks. In particular, extrapolating our result we have shown we can recover previous calculations with massless quarks. Their application to LEP data, together with the already known LO, can provide a new way for determining the bottom quark mass and to show for the first time its running. \vskip 5mm \noindent{\bf Acknowledgements.} I would like to thank J.~Fuster for very encouraging comments during the development of this calculation and for carefully reading this manuscript, A.~Santamaria for very useful discussions and S. Narison for the very kind atmosphere created at Montpellier.
proofpile-arXiv_065-444
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} The recently reported Bose-Einstein condensates of trapped neutral atoms \cite{AE} -- \cite{KM} represent the first unambiguous observations of a weakly interacting bose condensed gas. Quantitatively, we can characterize the strength of the interaction by the expansion parameter in the perturbation treatment of the homogeneous bose gas , $\sqrt{n a^{3}}$, where $n$ is the density and $a$ the scattering length of the inter-atomic potential. In the atomic-trap experiments, typically n $\sim 10^{12}$ $-$ $10^{14} \rm{cm}^{-3}$, and a $\sim 1-5 $ nm, so that $\sqrt{n a^{3}} \sim 10^{-2}$$\; - \; $ $3 \times 10^{-5}$. Thus, in the sense of perturbation theory, the observed condensates are indeed textbook examples of weakly interacting systems. For the uniform bose gas, perturbation theory leads to simple analytical results. Although the trapped condensates can be described by means of the Hartree-Fock Bogolubov equations \cite{GG}-\cite{HP}, the latter approach does not lend itself to an analytical perturbation treatment. Intuitively, one expects that a many-body system whose density varies slowly in space can be described locally as a homogeneous system. Based on this picture, the Thomas-Fermi method \cite{F} \cite{T} was proposed for the calculation of the electron density in a heavy atom. Lieb and Simon \cite{LS} showed that the treatment is exact in the limit when the atomic number goes to infinity. Application to a confined Bose condensate was pioneered by Goldman, Silvera and Legget \cite{GSL}, and recently reconsidered by Chou, Yang, and Yu \cite{CN}-\cite{CL}. As pointed out by Kagan, Shlyapnikov, and Walraven\cite{KSW}, the local-density description is valid when \begin{equation} \mu/\hbar\omega\gg 1 \; , \end{equation} where $\mu$ is the mean-field energy per particle, or chemical potential, and $\hbar\omega$ is the zeropoint energy in the trap. In this paper, we derive such a description from first principles within the framework of the variational technique. We emphasize that, unlike the practice of neglecting the kinetic energy term in the Gross-Pitaevski equation which in the recent literature is sometimes called the Thomas-Fermi approximation, the resulting variational description is not limited to the condensate, but describes the depletion, pressure and all other thermodynamic quantities. Furthermore, like the uniform gas, the Thomas-Fermi theory leads to a perturbation treatment of the weakly interacting condensates, giving simple analytical expressions for these quantities. Another important advantage of the Thomas Fermi treatment is that it can be generalized to describe finite temperature systems, as we shall discuss in future work. In this paper we focus on the bose gas at zero temperature. The paper is organized as follows. In section 2, we generalize the usual Bogolubov transformation to describe spatially inhomogeneous condensates. In section 3, we introduce the Wigner representation and gradient expansion, which provide the tools to describe the nearly homogeneous systems and make the Thomas-Fermi approximation. The advantage of this systematic approach to the Thomas-Fermi approximation is that it provides an estimate of the error incurred by the inhomogeneity of the condensate, allowing one to estimate the accuracy of the Thomas Fermi results. We consider this point to be very important in view of the fact that some traps, depending on the potential and the number of trapped atoms, are too far from homogeneity to be described by a Thomas-Fermi description. In addition, even if the Thomas-Fermi description is valid in the middle of the trap, it breaks down at the edge of the condensate. In sections 4 and 5, we obtain the mean-field description of the bose system in the Thomas-Fermi approximation. The equations, derived within the framework of the variational principle, provide a fully self-consistent description, indicating that the Thomas-Fermi decription is by no means limited to weakly interacting systems. This remark can be expected to be of future importance in the light of recent experimental efforts to obtain condensates of higher density. Nonetheless, because of the special interest in the weakly interacting systems, we proceed in section 6 to derive a perturbation treatment and obtain analytical results for quantities such as the chemical potential, the local depletion, pairing and pressure. With the experimental atomic-traps in mind, we apply the results of the general theory to the special case of a trapping potential that is of the type of a simple spherically symmetric harmonic oscillator in section 7. Finally, in section 8, we derive a density of states of the trapped weakly interacting condensate within the spirit of the Thomas-Fermi approximation. \section{Generalized Bogolubov Transformation} The Bogolubov quasi-particle concept \cite{NP} provides a very elegant description of the interacting Bose-Einstein condensate. The quasi-particles are represented by creation ($\eta^{\dagger}$) and annihilation ($\eta$) operators that are linear combinations of regular single-particle creation ($a^{\dagger}$) and annihilation ($a$) operators. In treating a homogeneous system, for which we can work in a basis of single-particle plane-wave states of momentum ${\bf k}$, the Bogolubov transformation which relates the quasi-particle and regular particular operators, takes on a particularly simple form, \begin{eqnarray} \eta^{\dagger}_{\bf k} &=& x_{\bf k} a^{\dagger}_{\bf k} + y_{\bf k} a_{-{\bf k}} \; , \nonumber \\ \nonumber \\ \eta_{\bf k} &=& x_{\bf k} a_{\bf k} + y_{\bf k} a^{\dagger}_{-{\bf k}} \; , \label{e:BT} \end{eqnarray} where for the purpose of describing the static properties of a condensate in equilibrium, we can limit the transformation parameters, $x_{\bf k},y_{\bf k}$ to real numbers. Furthermore, the isotropy of the many-body system suggests that the transformation parameters only depend on the magnitude of the momentum, $x_{\rm{\bf k}}$ = $x_{\rm{k}}$ and $y_{\rm{\bf k}}$ = $y_{\rm{k}}$. Requiring the quasi-particle operators to be canonical, $\left[ \eta_{\bf k},\eta^{\dagger}_{{\bf k}'} \right]= \delta_{{\bf k},{{\bf k}'}}, \left[ \eta_{\bf k},\eta_{\bf k'} \right] = \left[ \eta^{\dagger}_{\bf k}, \eta^{\dagger}_{{\bf k}'} \right] = 0$, gives an additional constraint to $x_{\rm{k}}$ and $y_{\rm{k}}$, \begin{equation} x^{2}_{k} - y^{2}_{k} = 1 , \label{e:cans} \end{equation} from which we can see that a single parameter $\sigma_{k}$, with $x_{k}$ = $\cosh\sigma_{k}$ and $y_{k}$ = $\sinh\sigma_{k}$, suffices to parametrize the Bogolubov transformation (\ref{e:BT}). In addition, with Eq.(\ref{e:cans}), we can also write the Bogolubov transformation as \begin{eqnarray} a^{\dagger}_{\bf k} &=& x_{k} \eta^{\dagger}_{\bf k} - y_{k} \eta_{-{\bf k}} \; , \nonumber \\ \nonumber \\ a_{\bf k} &=& x_{k} \eta_{\bf k} - y_{k} \eta^{\dagger}_{-{\bf k}} \; , \label{e:BTI} \end{eqnarray} which is the inverse transformation of (\ref{e:BT}). It is useful to define the following quantities, the ``distribution function'' $\rho$ and the ``pairing function'' $\Delta$: \begin{eqnarray} \rho_{\bf k} &=& \langle a_{\bf k}^{\dagger} a_{\bf k} \rangle = \frac{1}{2} \left[ \cosh 2 \sigma_{\bf k} -1 \right] \; , \nonumber\\ \Delta_{\bf k} &=& - \langle a_{\bf k} a_{-\bf k} \rangle = \frac{1}{2} \sinh 2 \sigma_{\bf k} \; , \end{eqnarray} where the brackets $\langle \; \rangle$ represent the ground state expectation value. The best values for $x_{\rm{k}}$ and $y_{\rm{k}}$ are obtained variationally by minimizing the ground state free energy. As stated, the above description (1)-(3) only applies to homogeneous systems, whereas the treatment of a general (inhomogeneous) condensate, as we shall show below, involves a Bogolubov transformation that is quite different in appearance, from the homogeneous case. However, we can expect the results of the homogeneous treatment to describe the `local' behavior of an inhomogeneous condensate, provided the spatial variations of the condensate are sufficiently slow. In describing many-particle Fermion systems, this intuitive picture forms one of the key ingredients of the well-known Thomas-Fermi description of slowly varying many-particle systems. To arrive at a general treatment, we choose to work with boson-field operators, $\hat{\Psi}({\bf x})$ and $\hat{\Psi}^\dagger({\bf x})$, an approach that offers the advantage of not having to specify a basis a priori. Furthermore, in the presence of a condensate, it is convenient to work with the fields $\hat{\psi}({\bf x})$ and $\hat{\psi}^{\dagger}({\bf x})$, which are displaced from the original fields $\hat{\Psi}({\bf x})$ and $\hat{\Psi}^{\dagger}({\bf x})$ by the expectation value $\phi({\bf x})$ of $\hat{\Psi}({\bf x})$, \begin{eqnarray} \hat{\Psi}({\bf x}) &=& \hat{\psi}({\bf x}) + \phi({\bf x}), \nonumber \\ \hat{\Psi}^{\dagger}({\bf x}) &=& \hat{\psi}^{\dagger}({\bf x}) + \phi^{\ast}({\bf x}), \end{eqnarray} where, for the purpose of describing the static properties of a condensate in equilibrium, $\phi$ can be taken to be real, and where $\hat{\psi}({\bf x})$ and $\hat{\psi}^{\dagger} ({\bf x})$ are the displaced field operators which satisfy the canonical commutation relation, $\left[ \hat{\psi}({\bf x}), \hat{\psi}^{\dagger} ({\bf x}') \right] $ $= \delta ({\bf x}-{\bf x}')$, and furthermore, \begin{equation} \langle \hat{\psi}({\bf x}) \rangle = \langle \hat{\psi}^{\dagger} ({\bf x}) \rangle = 0. \label{e:expo} \end{equation} We introduce the Bogolubov transformation as a general linear transformation relating the displaced fields to the quasi-particle fields, $\hat{\xi}({\bf x})$ and $\hat{\xi}^{\dagger}({\bf x})$, \begin{eqnarray} \hat{\psi}({\bf x}) &=& \int d^{3}z \left[X({\bf x},{\bf z})\hat{\xi}({\bf z}) -Y({\bf x},{\bf z})\hat{\xi}^{\dagger}({\bf z})\right] , \nonumber \\ \nonumber\\ \hat{\psi}^{\dagger}(\bf x) &=& \int d^{3}z \left[X^{\ast}({\bf x} ,{\bf z})\hat{\xi}^{\dagger}({\bf z}) -Y^{\ast}({\bf x},{\bf z})\hat{\xi}({\bf z})\right], \label{e:GBT} \end{eqnarray} which is the generalization of Eq.(\ref{e:cans}). The non-local nature of the generalized Bogolubov transformation (\ref{e:GBT}) should not be surprising $-$ the `homogeneous' Bogolubov transformation (\ref{e:BT}) can be written in the same form with the special feature, due to the homogeneity of the system, that $X({\bf x}, {\bf y})$ and $Y({\bf x}, {\bf y})$ only depend on ${\bf x} - {\bf y}$. Requiring the quasi-particle fields to be canonical, leads to \begin{equation} \int d^{3} z \left[ X({\bf x},{\bf z}) X({\bf z},{\bf y}) - Y({\bf x},{\bf z}) Y({\bf z},{\bf y})\right] = \delta({\bf x}-{\bf y}) \; , \label{e:cang} \end{equation} which is the generalization of (\ref{e:cans}). It is possible to derive equations for the inhomogeneous bose systems by variationally determining the best transformations $X$ and $Y$, minimizing the free energy. This however, is not the path we choose to follow here. Instead, we manipulate the generalized Bogolubov transformations in a manner similar to the procedure to obtain the Wigner distribution from the off-diagonal single-particle density function. Once this is achieved, the steps that lead to a Thomas-Fermi description are known from quantum transport theory. One interesting new aspect of this treatment is that the central object of the theory is not a distribution function, which in some sense can still be regarded as an observable, but a transformation. Although this transformation determines the value of all observables, it is clearly not an observable quantity by itself. \section{Wigner Representation and Gradient Expansion} Wigner showed that a quantum mechanical single-particle system, costumarily characterized by its wave function $ \Psi ({\bf x}) $, can alternatively be fully characterized by a different function, \begin{equation} \rho_{\rm{W}}({\bf R},{\bf p}) = \int d^{3} r \; \Psi^{\ast}({\bf R}+{\bf r}/2) \Psi({\bf R}- {\bf r}/2) \exp (i{\bf p}\cdot{\bf r}) , \label{e:WR} \end{equation} where here $-$ as in the rest of the paper $-$ we work in units in which $\hbar$=1. This function (\ref{e:WR}), known as the Wigner Distribution function, can be interpreted as a phase space distribution function \cite{WG} and leads to a description that is remarkably close to classical mechanics. The analogy with a classical phase space distribution function is not complete (for example $ \rho_{\rm W} $ can take on negative values), but can be justified by the fact that the quantum mechanical expectation value of observables are equal to the `phase space integrals' of the corresponding classical quantities, weighted by $(2 \pi)^{-3} \rho_{{\rm W}} $, \begin{eqnarray} \langle \Psi | f | \Psi \rangle & = & \int d^{3} x \; \Psi^{\ast}({\bf x}) f({\bf x}) \Psi({\bf x})\nonumber \\ & = & (2 \pi)^{-3} \int d^{3} p \; \int d^{3} R \; f({\bf R}) \; \rho_{\rm{W}}({\bf R},{\bf p}), \nonumber\\ \langle \Psi |\hat{{\bf p}} | \Psi \rangle & = & \int d^{3} x \; \Psi^{\ast}({\bf x}) \hat{{\bf p}} \Psi({\bf x}) \nonumber\\ & = & (2 \pi)^{-3} \int d^{3} p \; \int d^{3} R \; {\bf p} \; \rho_{\rm{W}}({\bf R},{\bf p}), \end{eqnarray} etc. More recently, the many-particle generalization of the Wigner distribution has found many important applications in diverse areas such as nuclear \cite{DL} and solid state physics \cite{DD}. An important motivation to work in the transformed representation of Eqs.(\ref{e:WR}), $({\bf x},{\bf x}') \rightarrow ({\bf R},{\bf p})$, \begin{equation} A_{W}({\bf R},{\bf p}) = \int d^{3} r \; A({\bf R} + {\bf r}/2, {\bf R} - {\bf r}/2) \exp (i{\bf p}\cdot{\bf r}) , \label{e:wrep} \end{equation} and its inverse \begin{equation} A({\bf x} , {\bf x}') = (2 \pi)^{-3} \int d^{3} p \; A_{W} (\left[ {\bf x}+{\bf x}' \right] /2,{\bf p}) \; \exp(-i{\bf p}\cdot\left[ {\bf x}-{\bf x}' \right]), \label{e:wrepi} \end{equation} which shall henceforth be referred to as the Wigner representation, is that it is extraordinarily well suited to describe nearly homogeneous systems. This convenient feature follows from the gradient expansion \cite{DL} -- \cite{DD}. The gradient expansion shows that, to first order, a `product' operator $C({\bf x},{\bf x}') = \int d^{3} z \; A({\bf x},{\bf z}) B({\bf z},{\bf x}') $, in the Wigner representation simply gives the algebraic product of A and B, $C_{W}({\bf R},{\bf p}) \approx A_{W}({\bf R},{\bf p}) B_{W}({\bf R},{\bf p})$. The higher-order corrections to this approximation can be written as a series of terms containing successively higher-order derivatives in the (${\bf R},{\bf p}$)$-$ coordinates, \begin{eqnarray} C_{W}({\bf R},{\bf p}) \approx A_{W}({\bf R},{\bf p}) B_{W}({\bf R},{\bf p}) + \frac{1}{2 \rm{i}} \sum_{j=1}^{3} \left[ \frac{\partial A_{W}}{\partial R_{j}} \frac{\partial B_{W}}{\partial p_{j}} - \frac{\partial A_{W}}{\partial p_{j}} \frac{\partial B_{W}}{\partial R_{j}} \right]\nonumber\\ - \frac{1}{8} \sum_{j=1}^{3} \left[ \frac{\partial^{2} A_{W}}{\partial R_{j}^{2}} \frac{\partial^{2} B_{W}}{\partial p_{j}^{2}} + \frac{\partial^{2} A_{W}}{\partial p_{j}^{2}} \frac{\partial^{2} B_{W}}{\partial R_{j}^{2}} - 2 \frac{\partial^{2} A_{W}}{\partial R_{j} \partial p_{j}} \frac{\partial^{2} B_{W}}{\partial R_{j} \partial p_{j}} \right] + \cdots \; . \label{e:gradexp} \end{eqnarray} The first order correction in the gradient expansion (\ref{e:gradexp}) is $\left\{ A_{W},B_{W} \right\}_{\rm{{PB}}}$, the Poisson bracket of $A_{W}$ and $B_{W}$. If we know that the range of $A_{W}$ and $B_{W}$ in ${\bf p}$-space is of the order of $p_{c}$, then the magnitude of the derivatives $\partial B_{W} / \partial p $ and $\partial^{2} B_{W} / \partial p^{2}$ in (\ref{e:gradexp}) can be estimated to be of the order of $B_{W}/p_{c}$ and $B_{W} /p_{c}^{2}$ respectively. This approximation will allow us to obtain a very simple estimate of the `inhomogeneity' error. At this point, we return to the generalized Bogolubov transformation, $X({\bf x},{\bf y}),Y({\bf x},{\bf y})$, of the previous section. Working in the Wigner representation and expanding the `canonicity' relation (\ref{e:cang}) between between $X$ and $Y$ in the manner of the gradient expansion, we find up to first order in the spatial derivatives, a relation that is similar to the constraint equation (\ref{e:cans}) of the homogeneous Bogolubov transformation, \begin{equation} X^{2}_{W}({\bf R},{\bf p}) - Y^{2}_{W}({\bf R},{\bf p}) \approx 1 . \label{e:canw} \end{equation} Consequently, the general Bogolubov transform can be parametrized in the same way as the Bogolubov transform for the homogeneous bose gas, $X_{W}({\bf R},{\bf p})$ = $\cosh[\sigma ({\bf R},{\bf p})]$, $Y_{W}({\bf R},{\bf p})$ = $\sinh[\sigma ({\bf R},{\bf p})]$, where for the slowly varying condensate, the $\sigma$ $-$ parameters depend on the momentum {\sl and} position : $\sigma ({\bf R},{\bf p})$. The distribution and pairing functions, $\rho({\bf{x}},{\bf{x}}')$ = $\langle \hat{\psi}^{\dagger} (\bf{x}) \hat{\psi} (\bf{x}') \rangle $ and $\Delta({\bf{x}},{\bf{x}}')$ = $ - \langle \hat{\psi} (\bf{x}) \hat{\psi} (\bf{x}') \rangle$ take on the following form in the Wigner representation: \begin{eqnarray} \rho_{W} ({\bf R},{\bf p}) &=& \frac{1}{2} \left[ \cosh (2 \sigma ({\bf R},{\bf p})) -1 \right] , \nonumber\\ \Delta_{W} ({\bf R},{\bf p}) &=& \frac{1}{2} \sinh (2 \sigma ({\bf R},{\bf p})) . \label{e:rds1} \end{eqnarray} The local $\sigma$ parametrization of the Bogolubov transformation is crucial to the Thomas-Fermi description and it is upon the validity of (\ref{e:canw}) that the Thomas-Fermi theory rests. The error introduced to (\ref{e:canw}) due to the inhomogeneity of the system can be estimated by the lowest order non-vanishing term in the gradient expansion (\ref{e:gradexp}). Notice that the first order term in the gradient expansion of (\ref{e:canw}) vanishes since it is the the sum of Poisson brackets of quantities with themselves. Consequently, the error has to be estimated from the second order term. \section{Energy Density} In the variational method, the quantity to minimize is F, the ground state free energy, which we can put in a `local' form, $F = \int d^{3} R \; f({\bf R})$, where $f({\bf R})$ is the energy density. We achieve this result in two steps. In the first step, we shift to the Wigner representation in the integrand for the mean field expression for the ground state energy. In the second step, we notice that the short-range nature of the inter-atomic interaction renders the resulting integrand essentially `local', i.e. the integrand contains only $\sl single$ (not double) integrals over the position variables. The ground state free energy is the expectation value of $\hat{H} - \mu \hat{N}$, where $\hat{H}$ is the many-body hamiltonian of the boson system, $\hat{N}$, the number operator and $\mu$, the chemical potential: \begin{equation} {\hat{H}} - \mu \hat{N} = \int d^3 x \; \hat{\Psi}({\bf x})^{\dag} \hat{h}({\bf x}) \hat{\Psi}({\bf x}) + \frac{1}{2}\int d^3 x \; d^3y \; \hat{\Psi}({\bf y})^{\dag}\hat{\Psi}({\bf x})^{\dag} V(|{\bf x}-{\bf y}|) \hat{\Psi}({\bf x}) \hat{\Psi}({\bf y}), \label{e:ham} \end{equation} where $V(|{\bf x}-{\bf y}|)$ represents the inter-atomic potential and $\hat{h}({\bf x})$ is the one-body part of the free energy, \begin{equation} \hat{h}({\bf x}) = -\frac{\nabla^{2}}{2 m} + V_{\rm ext}({\bf x}) -\mu, \end{equation} where $V_{\rm ext}({\bf x})$ is the external potential. The presence of a condensate displaces the field operators ${\hat{\Psi}} ({\bf x})$ by their expectation value $\phi ({\bf x})$. To generate the variational free energy, we shall use the mean field approximation, in which terms of first and third order in $\hat{\psi}$ and $\hat{\psi}^{\dagger}$ vanish (\ref{e:expo}) and the fourth order term factorizes as follows: \begin{eqnarray} && \langle \hat{\psi}({\bf y})^{\dag}\hat{\psi}({\bf x})^{\dag} \hat{\psi}({\bf x}) \hat{\psi}({\bf y}) \rangle \approx \nonumber\\ \nonumber\\ && \Delta ^{\ast}({\bf y},{\bf x}) \Delta ({\bf y},{\bf x}) + \rho ({\bf y},{\bf x}) \rho ({\bf x},{\bf y}) + \rho ({\bf x},{\bf x}) \rho ({\bf y},{\bf y}) . \end{eqnarray} The variational nature of this procedure is insured by the existence of a variational ground state that gives this type of factorization. In fact, the variational ground state corresponds to the choice of the gaussian wave functional \cite{HP}. The displacement of the fields and the factorization of the expectation values, although straightforward, gives rise to a somewhat lengthy expression for the free energy. It is then convenient to classify the different contributions by their order in $\phi$ and their functional dependence on $\rho$ and $\Delta$: \noindent 1. $h_{1}$ is the one-body contribution of zeroth order in $\phi$ to the ground state energy, \begin{equation} h_{1} = \frac{1}{2} \int d^{3}x \; d^{3}y \; \hat{h}({\bf x}) \rho ({\bf y},{\bf x}) \delta ({\bf x}-{\bf y}) . \end{equation} 2. In analogy with the Hartree-Fock theory, we call $ V_{\rm{dir}} $ given below, the direct energy contribution to the energy, \begin{equation} V_{\rm{dir}} = \frac{1}{2} \int d^{3}x \; d^{3}y \; \rho ({\bf y},{\bf y}) \rho ({\bf x},{\bf x}) V(|{\bf x}-{\bf y}|). \end{equation} 3. Using the same analogy to the Hartree-Fock treatment, the exchange energy, $V_{\rm{exch}} $, is equal to \begin{equation} V_{\rm{exch}} = \frac{1}{2} \int d^{3}x \; d^{3}y \; \rho ({\bf x},{\bf y}) \rho ({\bf y},{\bf x}) V(|{\bf x}-{\bf y}|). \end{equation} 4. Standard Hartree-Fock theory does not describe pairing and the pairing energy, $ V_{\rm pair} $, \begin{equation} V_{\rm{pair}} = \frac{1}{2} \int d^{3}x \; d^{3}y \; \Delta^{\ast} ({\bf y},{\bf x}) \Delta ({\bf y},{\bf x}) V(|{\bf x}-{\bf y}|) \; , \end{equation} is consequently absent from the Hartree-Fock expressions. In second order in $\phi$, we find contributions that can be obtained from the above terms by replacing either $\Delta ({\bf x},{\bf y})$ or $\rho ({\bf x},{\bf y})$ by $\phi ({\bf x}) \phi ({\bf y})$. \noindent 5. For example, the one-body contribution, due to the kinetic and potential energy of the condensate is $h_{1}^{\phi}$, where \begin{equation} h_{1}^{\phi} = \int d^{3}x \; \phi ({\bf x}) \hat{h}({\bf x}) \phi ({\bf x}) . \end{equation} 6. $ V_{\rm{dir}}^{\phi} $ is the direct contribution to the interaction energy, stemming from the interaction of the condensate with the particles that have been `forced' out of the condensate (depletion), \begin{equation} V_{\rm{dir}}^{\phi} = \frac{1}{2} \int d^{3}x \; d^{3}y \; \phi ({\bf y}) \phi ({\bf y}) \rho ({\bf x},{\bf x}) V(|{\bf x}-{\bf y}|). \end{equation} 7. Similarly, $ V_{\rm{exch}}^{\phi} $ is the exchange contribution of second order in $\phi$, \begin{equation} V_{\rm{exch}}^{\phi} = \frac{1}{2} \int d^{3}x \; d^{3}y \; \phi ({\bf y}) \phi ({\bf x}) \rho ({\bf y},{\bf x}) V(|{\bf x}-{\bf y}|). \end{equation} 8. We represent the pairing energy of the condensate with the particles out of the condensate by $ V_{\rm{pair}}^{\phi} $, \begin{equation} V_{\rm{pair}}^{\phi} = \frac{1}{2} \int d^{3}x \; d^{3}y \; \phi ({\bf y}) \phi ({\bf x}) \Delta ({\bf y},{\bf x}) V(|{\bf x}-{\bf y}|). \end{equation} 9. Finally, we denote the contribution of fourth order in $\phi$, representing the interaction energy of the condensate with itself, by $ V^{\phi\phi} $ : \begin{equation} V^{\phi\phi} = \frac{1}{2} \int d^{3}x \; d^{3}y \; \phi^{2} ({\bf y}) \phi^{2} ({\bf x}) V(|{\bf x}-{\bf y}|). \end{equation} With this notation, the mean-field expression for the ground state energy reads \begin{eqnarray} F &=& \langle \hat{H} - \mu \hat{N} \rangle \nonumber\\ &=& h_{1} + V_{\rm{dir}} + V_{\rm{exch}} + V_{\rm{pair}} + \nonumber\\ && h_{1}^{\phi} + 2 V_{\rm{dir}}^{\phi} + 2 V_{\rm{exch}}^{\phi} - 2 V_{\rm{pair}}^{\phi} + V^{\phi\phi} , \end{eqnarray} where the minus sign of the $V_{pair}^{\phi}$ term stems from the definition of $\Delta$ = $- \langle \hat{\psi} \hat{\psi} \rangle$. At this point, we introduce the Wigner representation into the integrands of the above contributions to the mean-field expressions for the ground state free energy. The resulting expressions resemble the corresponding terms for the homogeneous gas, with an additional label ${\bf R}$ over which is integrated. For the sake of notational convenience we introduce the following integration symbol $\int_{\bf R}$ or $\int_{\bf p}$, which represents the usual integral over all of space, $\int d^{3} {\bf R}$, if ${\bf R}$ is a position variable or $(2\pi)^{-3} \int d^{3} p $, if ${\bf p}$ is a momentum variable: \begin{eqnarray} \int_{\bf p} &\equiv& (2\pi)^{-3} \; \int d^{3} p \; , \nonumber\\ \int_{\bf R} &\equiv& \int d^{3} R \; . \end{eqnarray} The terms of zero order in $\phi$ then give \begin{eqnarray} h_{1} &=& \int_{\bf R} \; \int_{\bf p} \left[ \frac{p^{2}}{2m} + V({\bf R}) - \mu \right] \rho_{W} ({\bf R},{\bf p}), \nonumber\\ \nonumber\\ V_{\rm{exch}} &=& \int_{\bf R} \; \int_{\bf p} \; \int_{{\bf p}'} \; \rho_{W} ({\bf R},{\bf p}) v({\bf p}-{\bf p}') \rho_{W} ({\bf R},{\bf p}') \; ,\nonumber\\ V_{\rm{pair}} &=& \int_{\bf R} \; \int_{\bf p} \; \int_{{\bf p}'} \; \Delta_{W} ({\bf R},{\bf p}) v({\bf p}-{\bf p}') \Delta_{W} ({\bf R},{\bf p}') \; , \nonumber\\ V_{\rm{dir}} &=& \int_{\bf R} \; \int_{\bf r} \; \int_{\bf p} \; \int_{{\bf p}'} \; \int_{\bf q} \rho_{W} ({\bf R} - {\bf r}/2 ,{\bf p} ) \rho_{W} ({\bf R} + {\bf r}/2 ,{\bf p}') \exp (i {\bf q}\cdot{\bf r}) v({\bf q})\; , \end{eqnarray} where $v$ is the Fourier transform of the interaction potential, $v({\bf q})$ = $\int d^{3} r V({\bf r}) \exp (-i{\bf q}\cdot{\bf r})$. The terms that are of second order in $\phi$ can be obtained by replacing one $\rho$ or $\Delta$ by $\phi \phi$. In the Wigner representation, this procedure yields expressions that are similar to the corresponding terms of zero order in $\phi$ with $\rho_{W} ({\bf R},{\bf p})$ or $\Delta_{W} ({\bf R},{\bf p})$ replaced by a function $Q_{W}({\bf R},{\bf p})$, where \begin{equation} Q_{W}({\bf R},{\bf p}) = \int_{\bf r} \; \phi ({\bf R} + {\bf r}/2) \phi ({\bf R} - {\bf r}/2) \exp (i{\bf p}\cdot{\bf r}) . \end{equation} Notice that the contributions of second order in $\phi$, are non-local in the sense that their expressions contain integrals over more than one position variable. Nevertheless, if we consider the scale on which the physical quantities vary in space, or in momentum space, it becomes apparent that the non-local integrals can be approximated by local expressions. We illustrate this point by considering the exchange ($V_{exch}^{\phi}$) and pairing ($V_{pair}^{\phi}$) energies. The key to obtain local expressions is to notice that $Q_{W}({\bf R},{\bf p})$ varies with respect to ${\bf p}$ on the scale of ${\bf R}_{0}^{-1}$, where ${\bf R}_{0}$ is the size of the condensate. On the other hand, $v({\bf p}-{\bf p}')$ varies on the scale of $l_{ r} ^{-1}$ where $l_{ r}$ is the range of the atom-atom interaction. Typically ${\bf R}_{0} \gg l_{ r}$ so that $Q_{W}({\bf R},{\bf p})$ varies much more rapidly with respect to ${\bf p}$ than $v({\bf p}-{\bf p}')$. In fact, when ${\bf p}$ is large enough to make $v({\bf p}-{\bf p}')$ significantly different from $v({\bf p}')$, $Q_{W}({\bf R},{\bf p}) \approx 0$. Thus, we can replace $v({\bf p}-{\bf p}')$ by $v({\bf p}')$ in the integrands : \begin{eqnarray} V_{\rm{exch}}^{\phi} &\approx& \frac{1}{2} \int_{\bf R} \int_{\bf r} \int_{\bf p} \int_{{\bf p}'} \; \phi ({\bf R} + {\bf r}/2) \phi ({\bf R} - {\bf r}/2) \exp (i{\bf p}\cdot{\bf r}) v({\bf p}') \rho_{W} ({\bf R},{\bf p}') \nonumber\\ &=& \frac{1}{2} \phi^{2} ({\bf R}) \int_{\bf R} \int_{{\bf p}'} v({\bf p}') \rho_{W} ({\bf R},{\bf p}'), \nonumber\\ V_{\rm{pair}}^{\phi} &\approx& \frac{1}{2} \phi^{2} ({\bf R}) \int_{\bf R} \int_{{\bf p}'} \; v({\bf p}') \rho_{W} ({\bf R},{\bf p}') . \end{eqnarray} The same considerations regarding the relative magnitude of the relevant length scales show that we can similarly simplify the expression of the $\phi^{4}$ interaction energy, $V^{\phi\phi} $, and the direct interaction energies, $V_{\rm{dir}}$ and $V_{\rm{dir}}^{\phi} $. The local expressions are most easily obtained by considering the difference in length scales before introducing the Wigner representation. In coordinate space, we notice that $\rho ({\bf x},{\bf x}) \approx \rho ({\bf y},{\bf y})$ if $|{\bf x}-{\bf y}| \le l_{ r}$. Thus, we can replace $\rho ({\bf x},{\bf x})$ by $\rho ({\bf y},{\bf y})$ in an integrand if it is accompanied by $V(|{\bf x}-{\bf y}|)$: \begin{eqnarray} V_{\rm{dir}} &\approx& \frac{1}{2}\int d^{3} x \; d^{3} y \; \rho^{2} ({\bf x},{\bf x}) V(|{\bf x}-{\bf y}|) \nonumber\\ &=& \frac{1}{2} v(0) \int_{\bf R} \int_{\bf p} \; \int_{{\bf p}'} \; \rho_{W} ({\bf R},{\bf p}) \rho_{W} ({\bf R},{\bf p}') \; , \nonumber\\ V_{\rm{dir}}^{\phi} &\approx& \frac{1}{2} v(0) \int_{\bf R} \int_{\bf p} \; \phi^{2} ({\bf R}) \rho_{W} ({\bf R},{\bf p}) \nonumber\\ V^{\phi\phi} &\approx& \frac{1}{2} v(0) \int_{\bf R} \phi^{4} ({\bf R}) . \end{eqnarray} To conclude this section, we summarize the results by remarking that the Wigner representation and the length scale considerations bring the free energy in an almost-local form. We need to qualify that statement because of the appearance of the Laplacian, a non-local operator, in the $ h^{\phi} $ -contribution to the energy. In fact, it is the non-locality of this term that gives rise to a generalized Gross-Pitaevski or non-linear Schrodinger equation (NLSE). The resulting (almost-local) ground state free energy is $F = \int d^{3} R \; f({\bf R})$, where \begin{eqnarray} f({\bf R}) &=& \int_{\bf p} \; \left[ \frac{p^{2}}{2m} + V_{\rm{ext}} ({\bf R}) - \mu \right] \rho({\bf R},{\bf p}) + v_{\rm{exch}} ({\bf R}) + v_{\rm{dir}} ({\bf R}) + v_{\rm{pair}} ({\bf R}) \nonumber\\ & &+ \phi ({\bf R}) \left[ \frac{-\nabla ^{2}}{2m} + V_{\rm{ext}} ({\bf R}) - \mu \right] \phi ({\bf R}) + 2 v_{\rm{exch}}^{\phi} ({\bf R}) + 2 v_{\rm{dir}}^{\phi} ({\bf R}) - 2 v_{\rm{pair}}^{\phi} ({\bf R}) \nonumber\\ & & + \frac{1}{2} v(0) \phi^{4} ({\bf R}), \end{eqnarray} where the exchange, direct and pairing energy densities are the integrands of the corresponding interaction energy contributions to the free energy : \begin{eqnarray} v_{\rm{exch}} ({\bf R}) &=& \frac{1}{2} \int_{\bf p} \; \int_{{\bf p}'} \; \rho ({\bf R},{\bf p}) v({\bf p}-{\bf p}') \rho_{W} ({\bf R},{\bf p}') \; , \nonumber\\ v_{\rm{pair}}({\bf R}) &=& \frac{1}{2} \int_{\bf p} \; \int_{{\bf p}'} \; \Delta_{W} ({\bf R},{\bf p}) v({\bf p}-{\bf p}') \Delta_{W} ({\bf R},{\bf p}') \; , \nonumber\\ v_{\rm{dir}} ({\bf R}) &=& \frac{1}{2} v(0) \int_{\bf p} \; \int_{{\bf p}'} \; \rho_{W} ({\bf R},{\bf p}) \rho_{W} ({\bf R},{\bf p}') \; , \nonumber\\ v_{\rm{exch}}^{\phi} ({\bf R}) &=& \frac{1}{2} \phi^{2} ({\bf R}) \int_{\bf p} \; \rho_{W} ({\bf R},{\bf p}) v({\bf p}) \; , \nonumber\\ v_{\rm{pair}}^{\phi}({\bf R}) &=& \frac{1}{2} \phi^{2} ({\bf R}) \int_{\bf p} \; \Delta_{W} ({\bf R},{\bf p}) v({\bf p}) \; , \nonumber\\ v_{\rm{dir}}^{\phi} ({\bf R}) &=& \frac{1}{2} \phi^{2} ({\bf R}) v(0) \int_{\bf p} \; \rho_{W} ({\bf R},{\bf p}) \; . \end{eqnarray} Notice that the free energy and free energy density are functionals of $\Delta ({\bf R},{\bf p})$, $\rho ({\bf R},{\bf p})$ and $\phi ({\bf R})$. In the next section we determine the equilibrium values of $\Delta ({\bf R},{\bf p})$, $\rho ({\bf R},{\bf p})$ and $\phi ({\bf R})$ by minimizing $F\left[ \rho, \Delta, \phi ; \mu \right]$. \section{Self-Consistent Mean Field Theory} In this section we derive the self-consistent mean-field equations that describe the nearly-uniform bose condensate at zero temperature. In the variational method, one minimizes the mean-field ground state free energy $F\left[ \rho , \Delta , \phi ; \mu \right]$. Writing the integrands of the different contributions to the mean-field free energy in the Wigner representation, followed by the length scale arguments of the previous section showed that $F\left[ \rho_{W}, \Delta_{W}, \phi ; \mu \right]$ is essentially a local quantity. Finally, in the Thomas-Fermi limit of a nearly-homogeneous system, $\rho_{W} ({\bf R},{\bf p})$ and $\Delta_{W} ({\bf R},{\bf p})$ are parametrized by a single Bogolubov transformation parameter $\sigma ({\bf R},{\bf p})$ in the manner of Eq.(\ref{e:rds1}). Thus, to describe a nearly-homogeneous system, we minimize the Thomas-Fermi ground state free energy, which is obtained from the mean-field free energy, assuming that $\rho_{W}$ and $\Delta_{W}$ are parametrized by $\sigma$ (\ref{e:rds1}), $F \left[ \sigma, \phi ; \mu \right]$ $=$ $F\left[ \rho_{W}(\sigma), \Delta_{W}(\sigma), \phi ; \mu \right]$. We obtain the condensate wave function $\phi_{0} ({\bf R})$ and Bogolubov parameter $\sigma_{0} ({\bf R},{\bf p})$ that describe the condensate by varying $\sigma$ and $\phi$ independently to get a minimum in F : \begin{eqnarray} \left. \frac{\delta F }{\delta \phi ({\bf R})} \right|_{\sigma = \sigma_{0},\phi=\phi_{0}} &=& 0 \; , \; \; \; \; (\rm{NLSE})\; \nonumber \\ \left. \frac{\delta F } {\delta \sigma ({\bf R},{\bf p})} \right|_{\sigma = \sigma_{0},\phi=\phi_{0}} &=& 0. \label{e:sfe} \end{eqnarray} The $\phi$ variation, $\delta F / \delta \phi =0$, gives the non-linear Schrodinger Equation (NLSE). The $\sigma$ variation, $\delta F / \delta \sigma = 0, $ gives an equation for $\sigma_{0} ({\bf R},{\bf p})$. From $ \rho = \frac{1}{2} \left[ \cosh (2 \sigma) -1 \right]$ and $\Delta = \frac{1}{2} \sinh (2 \sigma)$ (\ref{e:rds1}), we find that $\partial \rho / \partial \sigma = \sinh (2 \sigma)$ and $\partial \Delta / \partial \sigma = \cosh (2 \sigma)$, so that $\delta F / \delta \sigma = 0$ is equivalent to \begin{equation} \tanh (2 {\sigma}_{0}) = \frac{ - \delta F / \delta \Delta_{W}} {\delta F / \delta \rho_{W}} . \label{e:tan} \end{equation} Now, several terms of the NLSE, as well as the functional derivatives $\delta F / \delta \Delta$ and $\delta F / \delta \rho$ (\ref{e:tan}), depend on $\sigma_{0}$ and $\phi_{0}$ so that the resulting equations have to be solved self-consistently. To make the self-consistent nature of the equations more explicit, we consider the $\sigma$-dependent contributions to the functional derivatives, $\delta V_{\rm{exch}} / \delta \rho$, $\delta V_{\rm{dir}} / \delta \rho$ and $\delta V_{\rm{pair}} / \delta \Delta$, which we shall call the generalized potentials, \begin{eqnarray} U_{\rm{exch}} ({\bf R},{\bf p}) &=& \delta V_{\rm{exch}} / \delta \rho_{W} ({\bf R},{\bf p}) \; \; = \int_{{\bf p}'} \; v({\bf p}-{\bf p}') \rho_{W} ({\bf R}, {\bf p}') , \nonumber\\ \nonumber\\ U_{\rm{dir}} ({\bf R}) &=& \delta V_{\rm{dir}} / \delta \rho_{W} ({\bf R},{\bf p}) \; \; = v(0) \int_{{\bf p}'} \; \rho_{W} ({\bf R}, {\bf p}') , \nonumber\\ \nonumber\\ U_{\rm{pair}} ({\bf R},{\bf p}) &=& \delta V_{\rm{pair}} / \delta \Delta_{W} ({\bf R},{\bf p}) \; \; = \int_{{\bf p}'} \; v({\bf p}-{\bf p}') \Delta_{W} ({\bf R}, {\bf p}') , \label{e:genp} \end{eqnarray} where we name the generalized potentials after the respective interaction energies of which they are the functional derivatives, $U_{\rm{exch}}$ is the exchange potential, $U_{\rm{dir}}$ the direct potential and $U_{\rm{pair}}$ the pairing potential. Writing the distribution and pairing function in the integrands of the generalized potentials in terms of $2\sigma$, we find with (\ref{e:tan}) that the generalized potentials implicitly depend on the functional derivatives of F : \begin{eqnarray} U_{\rm{exch}} ({\bf R},{\bf p}) &=& \int_{{\bf p}'} \; v({\bf p}-{\bf p}') \frac{1}{2} \left[ \frac{\delta F / \delta \rho} {\sqrt{(\delta F / \delta \rho)^{2} - (\delta F / \delta \Delta)^{2}}} -1 \right] , \nonumber\\ \nonumber\\ U_{\rm{dir}} ({\bf R}) &=& v(0) \int_{{\bf p}'} \; \frac{1}{2} \left[ \frac{\delta F / \delta \rho } {\sqrt{(\delta F / \delta \rho)^{2} - (\delta F / \delta \Delta)^{2}}} -1 \right] , \nonumber\\ \nonumber\\ U_{\rm{pair}} ({\bf R},{\bf p}) &=& \int_{{\bf p}'} \; v({\bf p}-{\bf p}') \; \frac{1}{2} \left[ \frac{ - \delta F / \delta \Delta} {\sqrt{(\delta F / \delta \rho)^{2} - (\delta F / \delta \Delta)^{2}}} \right], \label{e:genpo} \end{eqnarray} where it is understood that the functional derivatives in the integrands are evaluated at ${\bf R}$ and ${\bf p}'$. Functional differentiation shows that the functional derivatives of F in turn depend on the generalized potentials, \begin{eqnarray} \frac{\delta F}{\delta \Delta_{W} ({\bf R},{\bf p})} &=& U_{\rm{pair}} ({\bf R},{\bf p}) - \phi^{2} ({\bf R}) v({\bf p}) \; , \nonumber\\ \nonumber\\ \frac{\delta F}{\delta \rho_{W} ({\bf R},{\bf p})} &=& \frac{p^{2}}{2m} + V_{\rm{ext}} ({\bf R}) - \mu \nonumber\\ && + U_{\rm{exch}} ({\bf R},{\bf p}) + U_{\rm{dir}} ({\bf R},{\bf p}) + \phi^{2} ({\bf R}) \left[ v({\bf p})+v(0) \right] \; . \label{e:fund} \end{eqnarray} Thus, equations (\ref{e:genpo}) and (\ref{e:fund}) self-consistently determine the generalized potentials. Furthermore, there is a dependence on the condensate wave function $\phi$. The latter has to be obtained from the NLSE : \begin{equation} \left[ - \frac{\nabla^{2}}{2m} + V_{\rm{ext}} ({\bf R}) - \mu + U({\bf R}) + v(0) \phi^{2} ({\bf R}) \right] \phi({\bf R}) = 0 , \label{e:nlse} \end{equation} where the potential $U({\bf R})$, is equal to : \begin{equation} U({\bf R}) = U_{\rm{dir}} ({\bf R}) + U_{\rm{exch}} ({\bf R},0) - U_{\rm{pair}} ({\bf R},0) \; . \label{e:unlse} \end{equation} This potential term, which stems from the interaction of the condensate with the particles out of the condensate, is absent in the simplest (low density limit) form of the NLSE, usually encountered in the literature. The equations, (\ref{e:genpo}) (\ref{e:fund}) (\ref{e:nlse}) and (\ref{e:unlse}), are the full set of self-consistent mean field equations that describe the condensate in the Thomas-Fermi approximation. The self-consistent equations for the homogeneous gas \cite{GR}, are recovered by putting $V_{\rm{ext}} = 0$ and by assuming that $\phi$ is independent of position so that the kinetic energy contribution to the NLSE vanishes. Regarding the connection with the intuitive Thomas-Fermi model, we note that $\mu$ and $V_{\rm{ext}}$ in the self-consistent mean field equations always appear as $\mu - V_{\rm{ext}} ({\bf R})$, so that it is natural to define a local effective chemical potential: \begin{equation} \mu_{\rm{eff}} ({\bf R}) = \mu - V_{\rm{ext}} ({\bf R}). \label{e:mueff} \end{equation} In fact, this is the essence of the Thomas-Fermi description : the system is described locally as a homogeneous system with a position dependent effective chemical potential (\ref{e:mueff}). The solutions to the fully self-consistent equations determine the expectation value of all (static) physical observables as a function of the chemical potential $\mu$. One observable we can obtain in this manner is N, the number of trapped particles, \begin{equation} N(\mu) = \frac{\partial F}{\partial \mu} = \int_{\bf R} \int_{\bf p} \rho ({\bf R},{\bf p}) + \int_{\bf R} \; \phi^{2} ({\bf R}) , \end{equation} the inversion of which yields $\mu$(N), from which we can cast the results for the thermodynamic quantities in terms of the parameter that is controled or measured in the experiment $-$ the number of atoms N. \section{Low Density Limit} The self-consistent equations, (\ref{e:genpo}) (\ref{e:fund}) (\ref{e:nlse}) and (\ref{e:unlse}) can be solved iteratively. In the low density regime, where $\sqrt{n a^{3}} \ll 1$, we approximate the result by the expressions obtained after a single iteration, starting from $\sigma^{(0)}_{0} =0$ ($U^{(0)}_{\rm{exch}}= $ $U^{(0)}_{\rm{dir}} = $ $U^{(0)}_{\rm{pair}} = 0$, where the superscript indicates the order of the iteration). With this first guess we solve the NLSE and obtain the functional derivatives (\ref{e:genpo}), $\delta F/\delta \rho$, $\delta F/ \delta \Delta$, yielding the first-order $\sigma -$parameter (\ref{e:tan}), $\sigma^{(1)}$, and the generalized potentials (\ref{e:genp}), $U^{(1)}_{\rm{dir}}$, $U^{(1)}_{\rm{exch}}$, $U^{(1)}_{\rm{pair}}$. With these single iteration expressions we compute the expectation values of the observable quantities. In solving the NLSE, we shall assume that $\phi ({\bf R})$ varies slowly enough that we can also neglect the kinetic energy operator. To make the dependence on the scattering length explicit, we replace the potential by a pseudopotential, \begin{equation} V_{\rm{pseudo}} ({\bf r}) = \lambda \delta({\bf r}) \frac{\partial} {\partial r} r , \label{e:pseudo} \end{equation} where $\lambda = 4 \pi \hbar^{2} a /m$ and the derivative operator is necessary to remove the divergency in the ground state free energy \cite{HY}. Furthermore, we shall assume that $\phi ({\bf R})$ varies slowly enough that we can also neglect the kinetic energy operator in solving the NLSE (\ref{e:nlse}) : \begin{equation} \lambda \left[ \phi^{{(1)}} ({\bf R}) \right] ^{2} = \mu_{\rm{eff}} ({\bf R}) , \label{e:phio} \end{equation} where $\mu_{\rm{eff}}$ is the effective chemical potential (\ref{e:mueff}). The functional derivatives (\ref{e:fund}) are \begin{eqnarray} \frac{\delta F^{\rm{{(1)}}}}{\delta \Delta} &=& - \lambda \left[ \phi^{{(1)}} ({\bf R}) \right] ^{2} \; , \nonumber\\ \nonumber\\ \frac{\delta F^{\rm{{(1)}}}}{\delta \rho} &=& \frac{p^{2}}{2m} - \mu_{\rm{eff}} ({\bf R}) + 2 \lambda \left[ \phi^{{(1)}} ({\bf R}) \right] ^{2}. \end{eqnarray} Consequently, the single iteration value for the Bogolubov transformation parameter $\sigma$ is equal to \begin{eqnarray} \tanh \left[ 2 \sigma^{\rm{{(1)}}}_{0} ({\bf R},{\bf p}) \right] &=& \frac{ \lambda \phi^{2} ({\bf R})} {(p^2/2m) - \mu_{\rm{eff}} ({\bf R}) + 2 \lambda \phi^{2} ({\bf R})} \nonumber\\ \nonumber\\ &=& \frac{\mu_{\rm{eff}} ({\bf R})} {(p^2/2m) + \mu_{\rm{eff}} ({\bf R})} \; \; , \label{e:sigm1} \end{eqnarray} which can be recognized as the dilute uniform gas result if we put $\mu_{\rm{eff}} = \mu$. The expression for the Bogolubov parameter $\sigma^{{(1)}}_{0}$ from Eq.(\ref{e:sigm1}) is what we would have obtained with an effective energy density neglecting the interaction energies of the particles out of the condensate, $V_{\rm{dir}} $, $ V_{\rm{exch}} $ and $ V_{\rm{pair}} $. In other words, the effective ground state energy is $F_{\rm{eff}}$ = $\int d^{3} R \; f_{\rm{eff}} ({\bf R})$, where \begin{eqnarray} f_{\rm{eff}} ({\bf R}) &=& \int_{\bf p} \; \left[ \left[ \frac{p^{2}}{2m} - \mu_{\rm{eff}} ({\bf R}) + 2 \lambda \phi^{2} ({\bf R}) \right] \; \rho_{W} ({\bf R},{\bf p}) - \lambda \phi^{2} ({\bf R}) \Delta_{W} ({\bf R},{\bf p}) \right] \nonumber\\ && - \mu_{\rm{eff}} ({\bf R}) \phi^{2} ({\bf R}) + \frac{\lambda}{2} \phi^{4} ({\bf R}) . \label{e:feff} \end{eqnarray} We obtain the results for the observable quantities by calculating their expectation values from the single iteration $\sigma^{{(1)}}_{0}$ of Eq.(\ref{e:sigm1}). For example, the condensate wave function is determined from the NLSE : \begin{equation} \lambda \phi^{2} ({\bf R}) \approx \mu_{\rm{eff}} ({\bf R}) - U^{(1)} ({\bf R}), \label{e:gap1} \end{equation} where the potential $U({\bf R})$ is the sum of the generalized potentials at zero momentum (\ref{e:unlse}), \begin{equation} U^{(1)} ({\bf R}) = U^{(1)}_{\rm{exch}} ({\bf R},0) + U^{(1)}_{\rm{dir}} ({\bf R}) - U^{(1)}_{\rm{pair}} ({\bf R},0) , \nonumber\\ \end{equation} evaluated with the single-iteration value for $\sigma$. The single-iteration values for the generalized potentials are computed to be : \begin{eqnarray} U^{(1)}_{\rm{exch}} ({\bf R},0) &=& U^{(1)}_{\rm{dir}} ({\bf R}) = \frac{\lambda}{3\pi^2} \left[ \mu_{\rm{eff}} ({\bf R}) \right]^{3/2} m^{3/2} \; , \nonumber\\ U^{(1)}_{\rm{pair}} ({\bf R},0) &=& -{\lambda\over\pi^2} \; \left[ \mu_{\rm{eff}} ({\bf R}) \right] ^{3/2} m^{3/2} \; . \label{e:u1} \end{eqnarray} Thus, the condensate density is (\ref{e:gap1}) \begin{equation} \phi^{2} ({\bf R}) \approx {1\over\lambda}\mu_{\rm eff} ({\bf R}) - \frac{5}{3 \pi^{2}} \left[\mu_{\rm{eff}} ({\bf R}) \right] ^{3/2} m^{3/2}\; \; . \label{e:phi1} \end{equation} The total density n(${\bf R}$), including the correction to $\phi^{2} ({\bf R})$ (\ref{e:phi1}) and the local depletion, is equal to \begin{eqnarray} n({\bf R}) &=& \phi ^{2} ({\bf R}) + \; \int_{\bf p} \; \rho ({\bf R},{\bf p})\nonumber\\ &\approx& \phi ^{2} ({\bf R}) + \frac{1}{3\pi^2} \left[\mu_{\rm{eff}} ({\bf R}) \right] ^{3/2} m^{3/2}\nonumber\\ &\approx& {1\over\lambda}\mu_{\rm{eff}} ({\bf R}) - \frac{4}{3 \pi^{2}} \left[\mu_{\rm{eff}} ({\bf R}) \right] ^{3/2} m^{3/2} , \label{e:npert} \end{eqnarray} resulting in an expression for the density, n(${\bf R}$), in terms of the effective chemical potential $\mu_{\rm{eff}} ({\bf R})$. Inverting this relation up to first order in $\sqrt{n a^{3}}$, we obtain \begin{equation} \mu_{\rm{eff}} ({\bf R}) \approx \lambda n({\bf R}) \left[ 1 + \frac{32}{3} \sqrt{\frac{n({\bf R}) a^{3}}{\pi}} \right] , \label{e:mu1} \end{equation} which, for the homogeneous case, reduces to the well-known perturbation result. Finally, in a similar manner, we obtain the local pressure $P({\bf R})$ from the expression for the effective free energy density (\ref{e:feff}), $P({\bf R})$ = $ - f^{(1)}_{\rm{eff}} ({\bf R})$, \begin{equation} P({\bf R}) = \frac{\lambda \phi^{4} ({\bf R})}{2} \left[ 1 - \frac{128}{15\pi^2} \sqrt{ n({\bf R}) a^{3}} \right]. \label{e:pres} \end{equation} We can then replace $\phi^{2}$ in (\ref{e:pres}) by its single-iteration value (\ref{e:phi1}). Furthermore, replacing $\mu_{\rm{eff}} ({\bf R})$ in the resulting expression by (\ref{e:mu1}) results in a local equation of state. The above results illustrate an important advantage of the Thomas-Fermi description $-$ by neglecting the kinetic energy operator in the NLSE we recover simple analytical expressions for most quantities. These expressions are the analogues of the perturbation results for the dilute homogeneous bose gas. It is then of course very important to determine the regime and the conditions under which these results can be trusted. One source of error in the theory stems from neglecting the Laplacian operator in the NLSE. This approximation, although convenient, is not part of the Thomas-Fermi description. It is always possible to calculate the condensate wave function numerically from the NLSE and proceed from there with the iteration of the self-consistent Thomas-Fermi equations (\ref{e:sfe})! Nevertheless, if we omit the Laplacian term, we can estimate the error by calculating $e_{L}$, the ratio of the kinetic energy term, $-\nabla^{2} \phi /2m$, and the non-linear potential energy in the NLSE, $\lambda \phi^{3}$, \begin{equation} e_{L} ({\bf R}) = | -\nabla^{2} \phi / 2m \lambda \phi^{3} | = \left| \frac{ -\nabla^{2} \phi ({\bf R}) / \phi ({\bf R}) }{k_{c}^{2}({\bf R})} \right|, \label{e:el} \end{equation} where $k_{c}({\bf R})$ = $\left[ 8 \pi a n({\bf R}) \right] ^{\frac{1}{2}}$ is the inverse of the local coherence length, $k_{c}$ = $\lambda_{c}^{-1}({\bf R})$. Another source of error, which cannot be remedied but is truly inherent to the Thomas-Fermi approximation, stems from the inhomogeneity of the system. This error is also more difficult to estimate, and one benefit of our approach is that the gradient expansion offers a `handle' on this quantity. Indeed, we use the lowest order non-vanishing term in the gradient expansion to estimate the error. This term is of second order because the first order term vanishes. We estimate its magnitude (\ref{e:gradexp}) by replacing the partial derivatives with respect to the momentum variables by $k_{c}^{-1}$, since $k_{c}$ is a measure of the range in ${\bf p}$ of the observable at zero temperature. The relative error for the general product of two arbitrary operators A and B, $e_{\rm{i}} \left[ AB \right] $, is then given by \begin{equation} e_{\rm{i}} \left[ AB \right] \approx \frac{1}{8 k^{2}_{c} ({\bf R})} \left[ \frac{\nabla^{2} A_{W} }{ A_{W} } + \frac{ \nabla^{2} B_{W} }{ B_{W}} - 2 \frac{ \nabla A_{W}\; \cdot \nabla B_{W} }{ A_{W} B_{W} } \right] , \end{equation} The validity of the Thomas-Fermi description depends on $X_{W}^{2}-Y_{W}^{2}=1$ (\ref{e:canw}), so that we use the accuracy of this equality to test the validity of the local homogeneity description. The expression (\ref{e:canw}) can also be written as $\exp \left[ \sigma({\bf R},{\bf p}) \right] \exp \left[ -\sigma({\bf R},{\bf p}) \right] =1 $, so that we choose $A_{W}$ as $\exp \left[ \sigma({\bf R},{\bf p}) \right]$ and $B_{W}$ as $\exp \left[ -\sigma({\bf R},{\bf p}) \right]$ to estimate the relative error $e_{\rm i}$. In fact, it is more convenient to work with $\exp(4\sigma)$ then $\exp(\sigma)$, so that we compute the inhomogeneity error $e_{i} \left[ \exp(4\sigma) \exp(-4\sigma)\right] $ of $\exp(4\sigma)$ and divide by 4 (since the relative error of $f^{n}$ is simply n $\times$ the relative error of f). In this manner, we find that \begin{eqnarray} e_{i} \left[ \exp(\sigma) \exp(-\sigma) \right] &=& \frac{1}{4} e_{i} \left[ \exp(4\sigma) \exp(-4\sigma) \right] \nonumber\\ &\approx& \frac{1}{8 k_{c}^{2} ({\bf R})} \left| \frac{\nabla \exp(4\sigma)} {\exp(4\sigma)} \right| ^{2} \; . \label{e:inh} \end{eqnarray} With the single-iteration value for the low-density condensate, \begin{equation} \exp \left[ 4\sigma({\bf R},{\bf p}) \right] = 1 + \frac{2 \mu_{\rm{eff}} ({\bf R})}{p^{2}/2m} , \end{equation} we find that the inhomogeneity error, $e_{i} ({\bf R})$ (\ref{e:inh}), is equal to \begin{equation} e_{i} ({\bf R}) = \frac{1}{2} \left| \frac{{\bf F}_{\rm{ext}}({\bf R}) \lambda_{c} ({\bf R})} {(p^{2}/{2m}) + 2 \mu_{\rm{eff}} ({\bf R})} \right| ^{2} , \end{equation} where ${\bf F}_{\rm{ext}}$ is the force of the external potential, ${\bf F}_{\rm{ext}}$ = $- \nabla V_{ext}$. As expected, the error is largest for ${\bf p} = 0$, and using the ${\bf p} = 0$ $-$ value, we obtain a simple position dependent estimate for the inhomogeneity error, $e_{i} ({\bf R})$, \begin{equation} e_{i} ({\bf R}) = \frac{1}{8} |{\bf F}_{\rm{ext}}({\bf R}) \lambda_{c}({\bf R}) / \lambda \phi^{2} ({\bf R}) |^{2} , \label{e:inhf} \end{equation} where we replaced $\mu_{\rm{eff}}$ by $ \lambda \phi^{2}$. By equating this error (\ref{e:inhf}) to a chosen value, $e_{\rm{cut}} \ll 1$, reflecting the accuracy we demand from the theory, we can determine the spatial boundary beyond which the Thomas Fermi Theory is less accurate than $e_{\rm{cut}}$. \section{Spherically Symmetric Harmonic Oscillator Trap} We now specialize $V_{\rm{ext}} ({\bf R})$ to a harmonic oscillator potential, \begin{equation} V_{\rm ext} ({\bf R}) = \frac{1}{2} \hbar \omega (R/L)^2 \; , \end{equation} where $L$ is the size of the harmonic oscilator ground state, \begin{equation} L = \sqrt{\frac{\hbar}{m \omega}}, \end{equation} and compute the expectation value of important quantities in the low density limit of the previous section. In zeroth order in the iteration, we recover the results of Baym and Pethick \cite{GB}. From (\ref{e:phio}) we see that \begin{eqnarray} \left[ \phi^{(0)} ({\bf R}) \right] ^{2} &=& \left[ \mu - V_{\rm{ext}} ({\bf R}) \right] / \lambda \nonumber\\ &=& \frac{R^{2}_{0}}{8 \pi a L^{4}} \left[ 1 - (R/R_{0})^{2} \right] \label{e:phios} \end{eqnarray} where $R_{0}$ is the size of the condensate, $R_{0} = \sqrt{\frac{2 \mu}{\hbar \omega}} L$. In zeroth order, all particles are in the condensate, so that N $= \int_{\bf R} \phi^{2} ({\bf R}) $, and \begin{equation} \mu^{(0)} = \frac{\hbar \omega}{2} \left(\frac{ 15 a N}{L} \right) ^{2/5} , \end{equation} and consequently, \begin{equation} R_{0} = L \left( \frac{15 a N}{L} \right) ^{1/5} . \end{equation} The local coherence length, $\lambda_{c} ({\bf R})$ is given by \begin{equation} \lambda_{c} ({\bf R}) = \frac{L^{2}}{\sqrt{R_{0}^{2} - R^{2}}} \; \; . \label{e:cohl} \end{equation} Before we proceed to calculate the perturbation corrections to the observables, we test the validity of the low density Thomas-Fermi formalism by calculating the errors. The error due to neglecting the Laplacian in the NLSE, $e_{L} ({\bf R})$ (\ref{e:el}), is easily computed with (\ref{e:cohl}): \begin{equation} e_{L} ({\bf R}) = \left( \frac{L}{R_{0}} \right) ^{4} \frac{ \left[ 3 - 2 (R/R_{0})^{2} \right] }{(1 - (R/R_{0})^{2})^{3}} \; \; , \end{equation} from which we see that the laplacian can be omitted in the NLSE on condition that the size of the condensate is much larger than the size of the ground state, $ R_{0} \gg L $, or $ \left( 15 a N / L \right)^{1/5} \gg 1$. The error due to the departure of the BEC from homogeneity, $e_{i} ({\bf R})$ (\ref{e:inh}), is \begin{equation} e_{i} ({\bf R}) = \frac{1}{2} \left( \frac{L}{R_{0}} \right) ^{4} \frac{(R/R_{0})^{2}}{(1 - (R/R_{0})^{2})^{3}} \; \; . \end{equation} Again, notice that $e_{i}$ is small over most of the condensate region ($R < R_{0}$) if $ R_{0} \gg L $. In Fig.~(1) we show the density, $[\phi^{(0)} (R)]^{2} / [ \phi^{(0)} (R=0) ]^{2}$ (\ref{e:phios}), and both errors, $e_{L}$ and $e_{i}$, as a function of the distance to the middle of the trap. \begin{figure}[htbp] \centerline{\BoxedEPSF{fig1.eps scaled 600}} \caption{{\sf (a) Condensate density for N = $10^{3}$ and $10^{6}$; (b) Error incurred in neglecting kinetic term in NLSE; (c) Error incurred in Thomas-Fermi approximation. Length scale on horizontal axis is in units of L, the extend of the ground state wave function. Calculations are done for L = $10^{-4}$ cm, scattering length a = $5 \times 10^{-7}$ cm.}} \end{figure} The curves are calculated for a harmonic oscillator trap of $L = 1 \mu m$ and an inter-atomic interaction with scattering length a = 5 nm. The dotted lines correspond to $N = 10^{3}$ atoms in the trap, and the full line gives the results for $N = 10^{6}$ atoms. Notice that for $10^{3}$ particles, the Laplacian error is already substantial ($\sim 10 \%$) in the middle of the trap. In contrast, only at $ R = 1.8 L $ (to be compared to $ R_{0} = 2.4 L $) does the inhomogeneity error become of comparable magnitude. This indicates that even for as few as 1000 particles in the trap, the Thomas-Fermi description could be reasonably accurate for these parameters, provided one keeps the kinetic energy term in solving the NLSE. For $10^{6}$ atoms, $e_{i}$ and $e_{L}$ only become of the order of $ 10 \%$ at $ R = 9.0 L$, whereas $ R_{0} = 9.4 L$, which shows that the Thomas-Fermi description and neglecting the Laplacian operator are valid approximations in almost all of the condensate region. Under this condition, it is meaningful to calculate the perturbation corrections to the expectation values of the observable quantities. Including the perturbation correction, the local density (\ref{e:npert}) is equal to \begin{equation} n(R) = \frac{R_{0}^{2}}{8 \pi a L^{4}} \left[ 1 - (R/R_{0})^{2} \right] \left[ 1 - \frac{2 \sqrt{2}}{3 \pi} \frac{a R_{0}}{L^{2}} \sqrt{1 - \left(R/R_{0}\right)^{2} } \right]. \end{equation} The number of trapped particles, N, is obtained by integrating over the density $n({\bf R})$, \begin{equation} N = \int_{\bf R} \; n(R) = 4 \pi \int_{0}^{R_{0}} d R \; R^{2} \; n(R) \; , \label{e:nint} \end{equation} which leads to \begin{equation} N = \frac{1}{15} \frac{L}{a} \left(\frac{2 \mu}{\hbar \omega}\right)^{5/2} - \frac{\sqrt{2}}{24} \left(\frac{2 \mu}{\hbar \omega}\right)^{3} . \end{equation} The inverse relation, $\mu$ as a function of N, can be obtained by solving for $\mu$ iteratively in the previous equation, wich gives up to second iteration, the following result : \begin{equation} \mu = \frac{\hbar\omega}{2} \left(\frac{15 a}{L} \right)^{2/5} N^{2/5} \left[ 1 + \frac{\sqrt{2}}{60} \left(\frac{15 a}{L} \right)^{6/5} N^{1/5} \right]. \end{equation} Similarly, we obtain the condensate density from (\ref{e:phi1}) or the local depletion, $d({\bf R}) = \left[ n^{(1)} ({\bf R}) \right.$ $-\left. \left[ \phi^{1} ({\bf R}) \right] ^{2} \right] / \left[ \phi^{(1)} ({\bf R}) \right] ^{2} $ : \begin{equation} d(R) = \frac{2 \sqrt{2}}{3 \pi} \frac{a R_{0}}{L^{2}} \sqrt{ 1 - ( R/R_{0} )^{2} } . \end{equation} In Fig.~(2), we show the local depletion as a function of position for the same parameters as those of Fig.~(1). \medskip \begin{figure}[htbp] \centerline{\BoxedEPSF{fig2.eps scaled 600}} \caption{{\sf Depletion, defined as $d(R) = \left[ n({\bf R}) - \phi^{2} ({\bf R}) \right] / \phi^{2} ({\bf R})$, for the same systems as Fig.~(1).}} \end{figure} \medskip The local pressure is shown in Fig.~(3). \medskip \begin{figure}[htbp] \centerline{\BoxedEPSF{fig3.eps scaled 600}} \caption{{\sf Pressure for the same systems as Fig.~(1).}} \end{figure} \medskip To conclude this section, we repeat that the condition for the validity of the Thomas-Fermi description is that the size of the condensate exceeds the size of the ground state of the trap, $R_{0} \gg L$. An equivalent condition is that the coherence length in the middle of the condensate is smaller than the size of the ground state $\lambda_{c} (R=0) \ll L$, or that the chemical potential exceeds the ground state energy, $\mu \gg (\hbar \omega /2)$. These statements do not depend on the details of the trapping potential. Of course, the shape of the condensate, the boundary where the Thomas-Fermi description breaks down, and the expectation values of the local observables do depend on the shape of the potential. In this section, we gave the results for a spherically symmetric harmonic oscillator potential. For the convenience of the reader we tabulate several of the results up to first non-vanishing order in table I. \section{Density of States} In the Thomas-Fermi picture, the system is locally equivalent to a uniform system. Therefore, there are `local' excitations which in the low-density regime are described by the following energy spectrum : \begin{equation} {\epsilon}_{p} ({\bf R}) = \sqrt{(p^{2}/2m+\mu_{\rm{eff}}({\bf R}))^{2}-\mu_{\rm{eff}}^{2} ({\bf R}) } + \mu \; , \label{e:bspec} \end{equation} which is well known from the Bogolubov treatment of the uniform case. The local dispersion relation (\ref{e:bspec}) describes a phonon with position dependent sound velocity. To obtain the excitation of the whole system we compute the density of states using the formula \begin{equation} g(\epsilon) = \sum_{i} \; \delta(\epsilon - \epsilon_{i}), \label{e:dsc} \end{equation} where $\sum_{i}$ represents the sum over all excited states. In the spirit of the Thomas-Fermi approximation we take \begin{equation} g(\epsilon) = \int_{\bf R} \int_{\bf p} \; \delta \left(\epsilon - \epsilon_{\bf p} ({\bf R}) \right) \; . \label{e:dstf} \end{equation} After integration over the momentum variable, we obtain \begin{equation} g(\epsilon) = \frac{1}{2 \pi^{2}} \int_{\bf R} \; p_{\epsilon}^{2} ({\bf R}) \; \left| \frac{\partial \epsilon}{\partial p} \right| ^{-1} , \label{e:dss} \end{equation} where $p_{\epsilon} ({\bf R})$ is the momentum of a particle at position ${\bf R}$ with energy $\epsilon$. When calculating the remaining integral over space, we need to distinguish between spatial region (I) with condensate and a second region (II) without condensate, shown schematically in Fig.~(4). \medskip \begin{figure}[htbp] \centerline{\BoxedEPSF{fig4.eps scaled 600}} \caption{{\sf Schematic representation of the region with (region I), and without condensate (region II) for a BEC in a harmonic trap. The condensate density is proportional to $\mu_{\rm{eff}} ({\bf R})$, which is a `mirror image' of the trapping potential. Particles in the condensate have energy $\mu$ and a particle excited up to energy $\epsilon$ can move into region (II) as far as the classical turning point $R_{\epsilon}$.}} \end{figure} \medskip It is necessary to break up the integral (\ref{e:dss}) over the different integration regions, because the dispersion relations for the excitations are different. In region (I), we use the Bogoliubov spectrum (\ref{e:bspec}), whereas in region (II), the atoms are essentially free particles moving in the trap: \begin{equation} \epsilon_{p} ({\bf R}) = \frac{p^{2}}{2m} + V_{\rm{ext}} ({\bf R}) \; . \label{e:fspec} \end{equation} The density of states is then the sum of the integrals over region (I) and (II): \begin{eqnarray} g(\epsilon) = \frac{\sqrt{2}}{2} \frac{m^{3/2}}{\pi^{2}} & & \left[ (\epsilon - \mu) \int_{(I)} d^{3}R \; \frac{\sqrt{ \sqrt{\left[ \epsilon - \mu \right] ^{2} + \mu^{2}_{\rm{eff}}({\bf R})} - \mu_{\rm{eff}} ({\bf R})}} {\sqrt{\left[ \epsilon - \mu \right] ^{2} + \mu^{2}_{\rm{eff}}({\bf R})}} \right. \nonumber\\ && \left. + \int_{(II)} d^{3} R \; \sqrt{\epsilon - V_{\rm ext}({\bf R})} \right] . \end{eqnarray} For the special case of a spherically symmetric harmonic oscillator trap, we find the following expression for the density of states : \begin{eqnarray} g(\epsilon) = \frac{4}{\pi} \; \frac{\mu^{2}}{(\hbar \omega)^{3}} & & \left[ (\epsilon / \mu -1) \int_{0}^{1} dr \sqrt{1 - r} \frac{ \sqrt{ \sqrt{ (\epsilon / \mu -1)^{2} + r^{2}} - r }} { \sqrt{ (\epsilon / \mu -1)^{2} + r^{2}}} \right. \nonumber\\ & & \left. \; + \; 2 \int_{1}^{\epsilon / \mu} dr \; r^{2} \; \sqrt{ \epsilon / \mu - r^{2}} \; \right] \; . \end{eqnarray} In Figs. (5) and (6) we show the density of states for the system discussed in the previous section, L = $ 1 \mu m$, a = 5 nm, N = $10^{3}$ (Fig.~(5)) and N = $10^{6}$ (Fig.~ (6)). \medskip \begin{figure}[htbp] \centerline{\BoxedEPSF{fig5.eps scaled 600}} \caption{{\sf Density of states calculated in the Thomas-Fermi approach described in the paper. The system is a BEC of N = $10^{3}$ particles interacting with a scattering length a = $5 \times 10^{-7}$ cm, in a harmonic trap with ground state of extend L = $10^{-4}$ cm.}} \end{figure} \medskip \medskip \begin{figure}[htbp] \centerline{\BoxedEPSF{fig6.eps scaled 600}} \caption{{\sf Density of states for the same system as in Fig.~(6), but with N = $10^{6}$ particles.}} \end{figure} \medskip The dotted lines show the result for the interacting bose gas, the full line shows the density of states of the ideal gas in the same trap. The density of states starts from the chemical potential $\mu$, consistent with (\ref{e:bspec}), which implies that the energies are measured from the bottom of the potential well so that a particle of zero momentum in the condensate has energy $\mu$. If we were to set out the density of states as a function of excitation energy $\epsilon - \mu$, the density of states curves for the interacting BEC-systems would be shifted to the left by an amount $\mu$. In contrast to the homogeneous BEC, the density of states for the interacting case, as a function of the excitation energy, grows faster than the density of states of the ideal gas. The reason is purely geometrical: the phonon has a much larger volume in coordinate space available (at least the volume of the condensate) than the non-interacting boson that received the same amount of energy and can only move near the bottom of the potential well. This effect outweighs the fact that the momentum space volume available to the phonon is less than the momentum space volume available to the non-interacting particle with the same energy. Of course, the sharpness of the boundary between region (I) and (II), is an artifact of neglecting the Laplacian operator in the NLSE. Nevertheless, except for a region near the boundary, we argue that the rest of space is well-described and that the contribution of the near-boundary region is comparatively small so that the error that is introduced in the integral (\ref{e:dss}) is small provided the Thomas-Fermi description is valid in most of the condensate region. \section*{Acknowledgments} This work was supported in part by funds provided by the U.S. Department of Energy under cooperative agreement \# DE-FC02-94ER40818. P.T. was supported by Conselho Nacional de Desenvolvimento Cientifico e Tecnologico (CNPq), Brazil. The work of E.T. is supported by the NSF through a grant for the Institute for Atomic and Molecular Physics at Harvard University and Smithsonian Astrophysical Observatory. \newpage
proofpile-arXiv_065-445
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} Recently it was shown$^{1,2}$ that the Abrikosov-Gor'kov (AG)$^{3}$ theory of impure superconductors predicts a large decrease of $T_{c}$, linear in the nonmagnetic impurity concentration, which is not consistent with Anderson's theorem.$^{4}$ In their response$^{5}$ on ref. 1, AG argued that the frequency cutoff makes the AG theory compatible with Anderson's theorem. This argument is based on Tsuneto's application$^{6}$ of the AG theory to the Eliashberg equation. To settle this controversy, we first need to understand the limitation of Anderson's theorem. It was pointed out that Anderson's theorem is valid only up to the first order of the impurity concentration,$^{1,7}$ and the phonon-mediated interaction is strongly decreased by Anderson localization.$^{1,8,9}$ However, Tsuneto's theory fails to show the existence of localization correction to the phonon-mediated interaction. The failure comes from the intrinsic pairing problem$^{10,11}$ in Gor'kov's formalism.$^{12}$ The kernel of the self-consistency equation should be set by the physical constraint of the Anomalous Green's function. The resulting equation is nothing but another form of the BCS gap equation. Using the equation, the localization correction to the phonon-mediated interaction may be calculated.$^{1,11}$ A correct strong coupling theory has been reported by Kim.$^{8}$ For magnetic impurity effects, Kim and Overhauser (KO)$^{7}$ proposed a BCS type theory with different predictions: (i) The initial slope of $T_{c}$ decrease depends on the superconductor and is not the universal constant proposed by Abrikosov and Gor'kov(AG).$^{3}$ (ii) The $T_{c}$ reduction by exchange scattering is partially suppressed by potential scattering when the overall mean free path is smaller than the coherence length. This compensation has been confirmed in several experiments.$^{13-15}$ The difference comes again from the pairing problem. If we impose a correct pairing condition on the self-consistency equation, or the AG's calculation, we can find KO's result. \section{Phonon-Mediated Interaction in BCS Theory and Gor'kov's Formalism } \subsection{BCS Theory} For a homogeneous system, BCS introduced a reduced version of the Fr\"ohlich phonon-mediated interaction, \begin{eqnarray} H_{red} = \sum_{{\vec k} {\vec k}'} V_{{\vec k} {\vec k}'} c_{{\vec k}'}^{\dagger}c_{-{\vec k}'}^{\dagger} c_{-{\vec k}}c_{{\vec k}}, \end{eqnarray} where \begin{eqnarray} V_{{\vec k}{\vec k}'}= \cases{-V, &if $|\epsilon_{{\vec k}}|,|\epsilon_{{\vec k}'}|\leq \omega_{D}$\cr 0, &otherwise.\cr} \end{eqnarray} This reduction procedure is recognizing in advance which eigenstates will be paired and so contribute to the BCS condensate. In the presence of impurities, we can derive the phonon-mediated interaction by transforming the Fr\"ohlich interaction using the relation, \begin{eqnarray} \psi_{n\sigma} = \sum_{{\vec k}}\phi_{{\vec k}\sigma}<{\vec k}|n>. \end{eqnarray} $\psi_{n}$ and $\phi_{{\vec k}}$ denote the scattered state and the plane wave state. The reduced version of this interaction anticipates that $\psi_{n}$ (having spin up) will be paired with its time-reversed counterpart $\psi_{\overline {n}}$ (having spin down). The new reduced Hamiltonian is \begin{eqnarray} H_{red}' = \sum_{n n'} V_{n n'}c_{n'}^{\dagger}c_{\overline {n}'}^{\dagger}c_{\overline {n}}c_{n}, \end{eqnarray} where \begin{eqnarray} V_{n n'} = -V\sum_{{\vec k} {\vec k}' {\vec q}} <{\vec k} - {\vec q}|n'> <{\vec k}' + {\vec q}|\overline {n}'><{\vec k}'|\overline {n}>^{*}<{\vec k}|n>^{*}. \end{eqnarray} Anderson's theorem is valid only when ${\vec k}'$ can be set equal to $-{\vec k}$. \subsection{Gor'kov's formalism} In Gor'kov's formalism, a point interaction $-V\delta({\bf r}_{1}-{\bf r}_{2})$ is used for the pairing interaction between electrons. For a homogeneous system, the pairing interaction is \begin{eqnarray} H_{G} &=& - {1\over 2}V\int d{\bf r}\sum_{\alpha\beta}\Psi^{\dagger}({\bf r}\alpha) \Psi^{\dagger}({\bf r}\beta)\Psi({\bf r}\beta)\Psi({\bf r}\alpha) \nonumber \\ &=&-{1\over 2}V\sum_{{\vec k}{\vec k}'{\vec q}\sigma\sigma '} c_{{\vec k} - {\vec q}, \sigma}^{\dagger}c_{{\vec k}' + {\vec q}, \sigma '}^{\dagger}c_{{\vec k} '\sigma '} c_{{\vec k}, \sigma}, \end{eqnarray} and \begin{eqnarray} V_{{\vec k}{\vec k}'}&=&-V\int \phi_{{\vec k}'}^{*}({\bf r}) \phi_{-{\vec k}'}^{*}({\bf r}) \phi_{-{\vec k}}({\bf r}) \phi_{{\vec k}}({\bf r})d{\bf r}\nonumber \\ &=&-V. \end{eqnarray} Eq. (6) is the same as the Fr\"ohlich interaction within the BCS approximation. Note that the two points are not clear in Gor'kov's formalism, i.e., the BCS reduction procedure and the retardation cutoff. To obtain the same result as that of the BCS theory, these two ingredients should be taken care of in some way. As will be shown later, the negligence of the BCS reduction procedure causes a serious pairing problem especially in impure superconductors. In the presence of impurities, the matrix element of the pairing interaction is \begin{eqnarray} V_{nn'}=-V\int \psi_{n'}^{*}({\bf r}) \psi_{\bar n'}^{*}({\bf r}) \psi_{\bar n}({\bf r}) \psi_{n}({\bf r})d{\bf r}. \end{eqnarray} Substituting Eq. (3) into Eq. (8) we find that \begin{eqnarray} V_{nn'}=-V \sum_{{\vec k} {\vec k}' {\vec q}} <{\vec k} - {\vec q}|n'> <{\vec k}' + {\vec q}|\overline {n}'><{\vec k}'|\overline {n}>^{*}<{\vec k}|n>^{*}. \end{eqnarray} Notice that Eq. (9) is the same as Eq. (5). \section{Pairing Constraint on Gor'kov's Formalism} \subsection{Inhomogeneous System: Nonmagnetic Impurity Case} Near the transition temperature, the usual self-consistency equation is \begin{eqnarray} \Delta({\bf r}) &=& VT\sum_{\omega}\int \Delta({\bf l})G^{\uparrow}_{\omega}({\bf r,l}) G^{\downarrow}_{-\omega}({\bf r,l})d{\bf l}\nonumber \\ &=&\int K({\bf r},{{\bf l}})\Delta({{\bf l}})d{{\bf l}}. \end{eqnarray} Note that Kernel $K({\bf r},{{\bf l}})$ is not for Anderson's pairing. It includes the extra pairings between $n\uparrow$ and $n'(\not={\bar n})\downarrow$. The kernel for Anderson's pairing is \begin{eqnarray} K^{A}({\bf r},{{\bf l}})&=&VT\sum_{\omega} \{G^{\uparrow}_{\omega}({\bf r}, {{\bf l}}) G^{\downarrow}_{-\omega}({\bf r'}, {{\bf l}})\}_{p.p.}\nonumber \\ &=&V\sum_{n}{1\over 2\epsilon_{n}}tanh{\epsilon_{n}\over 2T}\psi_{n}({\bf r}) \psi_{\bar n}({\bf r}) \psi^{*}_{\bar n}({{\bf l}})\psi^{*}_{n}({{\bf l}}), \end{eqnarray} where p.p. means proper pairing constraint, which dictates pairing between $n\uparrow$ and ${\bar n}\downarrow$. It can be shown$^{10,11}$ that the extra pairings violate the physical constraint of the Anomalous Green's function, i.e., \begin{eqnarray} \overline{F({\bf r},{\bf r'},\omega)}^{imp}&\sim& \overline{\psi_{n\uparrow}({\bf r})\psi_{n'\downarrow}({\bf r'})}^{imp}\nonumber \\ &\not=& \overline{F({\bf r}-{\bf r'},\omega)}^{imp}. \end{eqnarray} These extra pairings should have been eliminated by the BCS reduction procedure in the Hamiltonian. Consequently, the revised self-consistency equation is \begin{eqnarray} \Delta({\bf r}) = VT\sum_{\omega}\int \Delta({{\bf l}})\{G^{\uparrow}_{\omega}({\bf r,{\bf l}}) G^{\downarrow}_{-\omega}({\bf r,{\bf l}})\}_{p.p.}d{\bf l}. \end{eqnarray} Notice that Eq. (13) is nothing but another form of the BCS gap equation, \begin{eqnarray} \Delta_{n}=\sum_{n'}V_{nn'}{\Delta_{n'}\over 2E_{n'}}tanh {E_{n'}\over 2T}. \end{eqnarray} \subsection{Inhomogeneous System: Magnetic Impurity Case} KO's $^{7}$ theory employed degenerate scattered state pairs. It has been claimed that the inclusion of the extra pairing is the origin of the so-called pair-breaking of the magnetic impurities.$^{16,17}$ However, the extra pairing terms cause the violation of the physical constraint of the pair potential and the Anomalous Green's function.$^{11}$ It can be shown$^{11}$ that the homogeneity condition of the Anomalous Green's function, after the impurity average, requires pairing between the degenerate scattered state partners. Then the revised self-consistency equation gives rise to KO's result.$^{7}$ \subsection{Homogeneous System} Near the transition temperature, the Anomalous Green's function is given by \begin{eqnarray} F({\bf r}, {\bf r'}, \omega) = \int \Delta({{\bf l}})G^{\uparrow}_{\omega}({\bf r}, {{\bf l}}) G^{\downarrow}_{-\omega}({\bf r'}, {{\bf l}})d{{\bf l}}, \end{eqnarray} Gor'kov$^{12}$ pointed out that $F({\bf r},{\bf r'})$ should depend only on ${\bf r}-{\bf r'}$, i.e., \begin{eqnarray} F({\bf r},{\bf r'}) = F({\bf r}-{\bf r'}). \end{eqnarray} Note that Eq. (15) includes the extra pairing terms between ${\vec k}\uparrow$ and ${\vec k}'\downarrow(\not= -{\vec k}\downarrow)$, which do not satisfy the homogeneity condition of Eq. (16). In this case, the self-consistency condition of the pair potential happens to eliminate the extra pairing in Eq. (15) because of the orthogonality of the wavefunctions. However, it is important to eliminate the extra pairing in the Anomalous Green's function from the beginning. Note that the kernel $K({\bf r},{{\bf l}})$ is not for the pairing between ${\vec k}\uparrow$ and $ -{\vec k}\downarrow$, but for the pairing between the states which are the linear combination of the plane wave states $\phi_{{\vec k}}({\bf r})$.$^{18}$ The inclusion of the extra pairings hindered our correct understanding of the impure superconductors and the relation between the pair potential and the gap parameter. \section{Pairing Constraint on the Bogoliubov-de Gennes Equations} \subsection{Inhomogeneous system: Nonmagnetic Impurity Case} By performing a unitary transformation, \begin{eqnarray} \Psi({\bf r}\uparrow) & =& \sum_{n}(\gamma_{n\uparrow}u_{n}({\bf r}) - \gamma^{\dagger}_{n\downarrow}v^{*}_{n}({\bf r})) \nonumber \\ \Psi({\bf r}\downarrow) & =& \sum_{n}(\gamma_{n\downarrow}u_{n}({\bf r}) + \gamma^{\dagger}_{n\uparrow}v^{*}_{n}({\bf r})), \end{eqnarray} we obtain the well-known Bogoliubov-de Gennes equations. To understand the physical meaning of the transformation (17), we express $\gamma_{n\uparrow}$ and $\gamma_{n\downarrow}$ by the creation and destruction operators for an electron in the scattered state,$^{18}$ \begin{eqnarray} \gamma_{n\uparrow} &= &\sum_{n'}\bigl( u_{n,n'}^{*}c_{n'\uparrow} + v_{n,n'}c_{n'\downarrow}^{\dagger}\bigr),\nonumber \\ \gamma_{n\downarrow} &= &\sum_{n'}\bigl( u_{n,n'}^{*}c_{n'\downarrow} - v_{n,n'}c_{n'\uparrow}^{\dagger}\bigr), \end{eqnarray} where \begin{eqnarray} u_{n,n'} & =& \int \psi^{*}_{n'}({\bf r}) u_{n}({\bf r})d{\bf r}\nonumber \\ v_{n,n'} & =& \int \psi^{*}_{n'}({\bf r})v^{*}_{n}({\bf r})d{\bf r}. \end{eqnarray} Accordingly, we obtain a vacuum state where $u_{n}({\bf r})\uparrow$ and $v_{n}^{*}({\bf r})\downarrow$ (instead of $\psi_{n}({\bf r})\uparrow$ and $ \psi_{\bar n}({\bf r})\downarrow$) are paired. They are the superpositions of the scattered states. It is clear that the Bogoliubov-de Gennes equations cannot give rise to the correct Anderson's pairing because of the position dependence of the pair potential. We need to supplement a pairing condition. If we assume a constant pair potential,$^{16,17}$ Anderson's pairing is obtained. However, then, the impurity effect on the phonon-mediated interaction is gone. In the presence of magnetic impurities, even the constant pair potential gives a pairing between the states which are the linear superpositions of the scattered states. \subsection{Homogeneous System } As in Sec. IV. A, the unitary transformation generates a vacuum state and the self-consistency equation for the pairing between $u_{n}({\bf r})\uparrow$ and $v^{*}_{n}({\bf r})\downarrow$, (instead of $\phi_{{\vec k}}({\bf r})\uparrow$ and $\phi_{-{\vec k}}({\bf r})\downarrow$). Note that $u_{n}({\bf r})\uparrow$ and $v^{*}_{n}({\bf r})\downarrow$ are the linear superpositions of the plane wave states until we constrain them. In this case, setting the pair potential gives a pairing between the plane wave states. However, the kernel of the self-consistency equation has not been set accordingly. \section{Pair Potential and Gap Parameter} For a homogeneous system, it was shown$^{19}$ \begin{eqnarray} \Delta({\bf r}-{\bf r'})=\int d{\vec k} e^{i{\vec k}\cdot({\bf r}-{\bf r'})}\Delta_{{\vec k}}. \end{eqnarray} But this relation is not exact because of the retardation cutoff. Correct relation may be obtained only after incorporating the pairing constraint into the self-consistency equation. It is given \begin{eqnarray} \Delta({\bf r}-{\bf r'})=V\sum_{{\vec k}}{\Delta_{{\vec k}}\over 2E_{{\vec k}}}tanh {E_{{\vec k}}\over 2T}\phi_{{\vec k}}({\bf r})\phi_{-{\vec k}}({\bf r'}). \end{eqnarray} Comparing Eq. (21) with the BCS gap equation and using Eq. (7), we also find \begin{eqnarray} \Delta_{{\vec k}}=\int \phi_{{\vec k}}^{*}({\bf r})\phi^{*}_{-{\vec k}}({\bf r}) \Delta({\bf r})d{\bf r}. \end{eqnarray} In the presence of impurities, one finds that \begin{eqnarray} \Delta({\bf r})=V\sum_{n}{\Delta_{n}\over 2E_{n}}tanh {E_{n}\over 2T}\psi_{n}({\bf r})\psi_{\bar n}({\bf r}), \end{eqnarray} and \begin{eqnarray} \Delta_{n}=\int \psi_{n}^{*}({\bf r})\psi^{*}_{\bar n}({\bf r}) \Delta({\bf r})d{\bf r}. \end{eqnarray} Eq. (24) was obtained first by Ma and Lee.$^{11}$ \section{Reinvestigation of Inhomogeneous Superconductors} Now we need to reinvestigate the inhomogeneous superconductors studied by Gor'kov's formalism or the Bogoliubov-de Gennes equations. In particular, Gor'kov's microscopic derivation$^{12}$ of the Gizburg-Landau equation is not valid. The gradient term may not be derived microscopically. To tackle the inhomogeneous problems, we must choose a correct pairing and calculate the correct kernel for each problem. Then the BCS theory may be more easy to apply. It is not clear whether the pair potential is the more appropriate quantity than the gap parameter in inhomogeneous superconductors. The following problems are required to restudy: 1. Dirty superconductors and localization 2. Magnetic impurities 3. Proximity effect, Andreev reflection and Josephson effect 4. Mesoscopic superconductivity 5. Type II superconductors, vortex problem 6. Non-equilibrium superconductivity 7. High $T_{c}$ superconductors, heavy fermion superconductors \subsection{Theory of Dirty Superconductors} Table I lists the theories of impure superconductors. Notice that Suhl and Matthias$^{20}$ and Abrikosov-Gor'kov$^{3}$ theories for $\Delta T_{c}$ are essentially equivalent. Both theories over-estimate the change of the density of states caused by impurity scattering, because they apply a retardation cutoff to the energy of the plane wave states and not of the scattered states. \vskip 4pt \centerline{{\bf TABLE I}. Theories of impure superconductors} \vskip 4pt \begin{tabular}{lll}\hline\hline & Ordinary impurity & Magnetic impurity\\ \hline Anderson & $T_{c} =T_{co}$ \qquad & \\ AG & $T_{c} = T_{co}-{T_{co}\over \pi\omega_{D}\tau}({1\over \lambda}+{1\over 2})$ \qquad & $T_{c}=T_{co}-{\pi\over 4}{1\over \tau_{s}}$\\ Suhl and Matthias & $T_{c}\cong T_{co}-{T_{co}\over \lambda\omega_{D}\tau}$ \qquad & $T_{c}=T_{co}-{\pi\over 3.5}{1\over \tau_{s}}$\\ Baltensperger & \qquad & $T_{c}=T_{co}-{\pi\over 4}{1\over \tau_{s}}$\\ Tsuneto & $T_{c}=T_{co}$ \qquad & \\ KO & $T_{c}=T_{co}-{T_{co}\over \pi\lambda E_{F}\tau}$ \qquad & $T_{c} =T_{co}-{0.18\pi\over \lambda\tau_{s}}$\cr & \qquad $-$ localization correction (Kim) \qquad & \\ \hline\hline \end{tabular} \vskip 8pt Now we consider strong coupling theories of dirty superconductors. Tsuneto$^{6}$ obtained the gap equation \begin{eqnarray} \Sigma_{2}(\omega) = {i\over (2\pi)^{3}p_{o}}\int dq\int d\epsilon \int d\omega ' {qD(q,\omega-\omega')\eta(\omega')\Sigma_{2}(\omega')\over \epsilon^{2} - \eta^{2}(\omega')\omega'^{2}}, \end{eqnarray} where $\eta=1 +{1\over 2\tau|\omega|}$, and $\tau$ is the collision time. On the other hand, Kim$^{8}$ obtained a gap equation \begin{eqnarray} \Delta^{*}(\omega_{n}, m) = \sum_{n'}\lambda(\omega_{n}-\omega_{n'}) \sum_{m'}V_{mm'}{\Delta^{*}(\omega_{n'},m')\over [-i\omega_{n'} -\epsilon_{m'}][i\omega_{n'}-\epsilon_{m'}]}, \end{eqnarray} where \begin{eqnarray} V_{mm'} = g^{2}\int |\psi_{m}({\bf r})|^{2} |\psi_{m'}({\bf r})|^{2}d{\bf r}. \end{eqnarray} Comparing Eqs. (25) and (26), we find that Tsuneto's result misses the most important factor $V_{mm'}$, which gives the change of the phonon-mediated interaction due to impurities. This factor is exponentially small for the localized states. \subsection{Suppression of Magnetic Impurity Effect by Ordinary Impurities} KO's results$^{7}$ can be explained physically. People used to notice that the impurity effect is stronger in the low $T_{c}$ material than in the high $T_{c}$ cuprate materials. Because the size of Cooper pair is very much smaller in high $T_{c}$ material, it sees a very small number of the impurities and so is not much influenced. Consequently, $T_{c}$ change and its initial slope depend on the material, which is predicted by KO. However, the AG theory predicts the universal slope. Because the AG theory pairs the states, which are the superpositions of the normal states of the material, the very nature of the material does not play an important role. When the conduction electrons have a mean free path that is smaller than the size of the Cooper pair (for a pure superconductor), the effective size is reduced. Accordingly, if we add ordinary impurities to the superconductors with the magnetic impurities enough to reduce the size of the Cooper pair, the magnetic impurity effect is partially suppressed. This compensation phenomena has been confirmed in several experiments.$^{13-15}$ The notion of gapless superconductor is based on the misunderstanding of the relation between the pair potential and the gap parameter. >From Eq. (23), it is clear that the pair potential cannot be finite when the gap parameter is zero. It seems that the long range order between the magnetic impurities or conduction electrons (especially in lead) caused the gaplesslike behavior in the experiments. Then it would rather be called zero gap superconductors or superconductors with nodes. \subsection{Weak Localization Correction to the Phonon-mediated Interaction} For the strongly localized states, the phonon-mediated interaction is exponentially small like the conductance. It is then expected that the same weak localization correction terms may occur in both quantities. Using the wavefunction obtained by Mott and Kaveh,$^{21}$ it can be shown that$^{11}$ \begin{eqnarray} V_{nn'}^{3d} &\cong& -V[1-{1\over (k_{F}\ell)^{2}}(1-{\ell\over L})],\nonumber \\ V_{nn'}^{2d} &\cong& -V[1-{2\over \pi k_{F}\ell}ln(L/\ell)],\nonumber \\ V_{nn'}^{1d} &\cong& -V[1-{1\over (\pi k_{F}a)^{2}}(L/\ell-1)], \end{eqnarray} where $a$ is the radius of the wire. There are many experimental results which show the reduction of $T_{c}$ caused by weak localization.$^{22,23}$ Previously, it was interpreted by the enhanced Coulomb repulsion. However, Dynes et al.$^{23}$ found a decrease of the Coulomb pseudo-potential $\mu^{*}$ with decreasing $T_{c}$. We believe that this signals the importance of weak localization correction to the phonon-mediated interaction. \subsection{Other Inhomogeneous Superconductors} We call attention to a few remarks against the conventional real space formalism. ``By the use of the Gor'kov technique, Abrikosov and Gor'kov have succeeded in $\cdots$, \quad it is {\sl entirely incorrect} as far as any physical results are concerned."$^{19}$ ``Certain {\sl inconsistencies} are seen to develop with the use of this approach (the de Gennes-Werthamer model), calling into question previous results obtained."$^{24}$ ``Is the discrepancy $\cdots$ indicative of the certain {\sl fundamental inconsistencies} in the Green's function formulation of the theory of nonstationary superconductivity?"$^{25}$ \section{Conclusions} It is shown that Gor'kov's formalism and the Bogoliubov-de Gennes equations need a pairing constraint. The resulting self-consistency equation is nothing but another form of the BCS gap equation. Most inhomogeneous superconductors should be reinvestigated. I am grateful to Professor A. W. Overhauser for discussions. This work was supported by the National Science Foundation, Materials Theory Program.
proofpile-arXiv_065-446
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section*{I. Introduction} \end{flushleft} The inclusive rare decay $B\rightarrow X_s\gamma$ has been studied several years before [1]. Recently the physics of $B_c$ meson has caught intensive attentions [2]. The $B_c$ meson is believed to be the next and the final family of $B$ mesons. It provides unique opportunity to examine various heavy quark fragmentation models, heavy quark spin-flavor symmetry, different quarkonium bound state models and properties of inclusive decay channels. Being made of two heavy quarks of different flavors, $B_c$ radiative weak decay also offer a rich source to measure elements of Cabibbo-Kobayashi- Maskawa ( CKM ) matrix of the standard model ( SM ). Different from the general rare $B$ decays $B\rightarrow X_s\gamma$ which is mainly induced by the flavor-changing $b\rightarrow s\gamma$ neutral currents [3], in $B_c\rightarrow D_s^*\gamma$, the bound state effects could seriously modify the results from the assumption. Bound state effects include modifications from weak annihilation which involve no neutral flavor-changing currents at all. The effects of weak annihilation mechanism are expected large due to the large CKM amplitude. Unfortunately, the well known chiral-symmetry [4] and the heavy quark symmetry [5] can not be applied to $B_c\rightarrow D_s^*\gamma$ process. Recently, a perturbative QCD ( PQCD ) analysis of $B$ meson decays seems to give a good prediction [6]. As it is argued in Ref.[7], $B_c$ two body nonleptonic decay can be conveniently studied within the framework of PQCD suggested by Brodsky-Lepage [8] and then developed in Ref.[6]. Here, we summarize their idea: In the subprocess $b\rightarrow s\gamma$, s quark obtains a large momentum by recoiling, in order to form a bound state with the spectator $\overline c$ quark, the most momentum of s must be transferred to $\overline c$ by a hard scattering process. In the final bound state ( i.e. $D_s^*$ ), since the heavy charm should share the most momentum of $D_s^*$, the hard scattering is suitable for PQCD calculation [6, 8]. In $B_c\rightarrow D_s^* \gamma$, the subprocess $b\rightarrow s\gamma$, taken as a free decay, is usually controlled by the one-loop electromagnetic penguin diagrams which are in particular sensitive to contributions from those new physics beyond the SM. This situation is similar with $B\rightarrow K^*\gamma$. Recently, the modification of the electromagnetic penguin interaction to the decay $b\rightarrow s\gamma$ from PGBs in the one generation technicolor model ( OGTM ) has been estimated in Ref.[9]. In the recent literatures [10, 11], the technicolor ( TC ) models with scalars have been studied extensively. The phenomenological studies have shown that the TC models with scalars do not generate unacceptably large FCNCs and are consistent with the experimental constraints on oblique electroweak radiative corrections. Among these models, the TC with a massless scalar doublet model ( TCMLSM ) [10], presented by C. D. Carone and H. Georgi is the simplest nontrivial extension of the SM with only two new free parameters ( $h_+$, $h_-$ or $h$, $\lambda$ ). The phenomenology of the TCMLSM has been discussed in the literatures [11]. In this paper we express these two new free parameters as $m_{\pi_p}$ and $\frac{f}{f'}$ differently from Ref. [10]. The relation of them can be found in Ref.[10], where $m_{\pi_p}$ is the mass of $\pi_p$ ( the physical pions in this model ), $f$ and $f'$ are the technipion decay constant and scalar vacuum expectation value ( VEV ), respectively. The couplings of charged $\pi_p$ with ordinary fermions are given as $$ [\pi_p^+--u_i--d_j]=-\frac{\sqrt2}{2}V_{u_id_j}\frac{f}{vf'}[m_{d_j}(1 +\gamma_5)-m_{u_i}(1-\gamma_5)], \eqno{(1)} $$ $$ [\pi_p^---u_i--d_j]=-\frac{\sqrt2}{2}V_{u_id_j}\frac{f}{vf'}[m_{u_i}(1 +\gamma_5)-m_{d_j}(1-\gamma_5)], \eqno{(2)} $$ where u=( u, c, t ), d=( d, s, b ), $V_{u_id_j}$ is the element of CKM matrix, and $v\sim$ 250 GeV is the electroweak scale. In this paper, applying the above PQCD method, we address $B_c\rightarrow D_s^*\gamma$ in the TCMLSM to examine the virtual effects of $\pi_p$ in the TCMLSM and compare the results with which estimated in the SM. This paper is organized as follows: In Sec.II, we display our calculations in the SM and TCMLSM. We present the final numerical results in Sec.III. Sec.IV contains the discussion. \begin{flushleft} \section*{II. Calculation} \end{flushleft} Using the factorization scheme [8] within PQCD, the momentum of quarks are taken as some fractions $x$ of the total momentum of the meson weighted by a soft physics distribution functions $\Phi_{H}(x)$. The meson wave functions of $B_c$ and $D_s^*$ take the simple form of $\delta$ function ( the so-called peaking approximation ) [12, 13]: $$ \Phi_{B_c}(x)=\frac{f_{B_c}\delta(x-\epsilon_{B_c})}{2\sqrt{3}}, \eqno{(3)} $$ $$ \Phi_{D_s^{*}}(y)=\frac{f_{D_s^{*}}\delta(y-\epsilon_{D_s^{*}})}{2\sqrt{3}}, \eqno{(4)} $$ The normalization [13] is $$ \int_{0}^{1} dx\Phi_{B_c}(x)=\frac{f_{B_c}}{2\sqrt{3}}, \eqno{(5)} $$ $$ \int_{0}^{1} dy\Phi_{D_s^*}(y)=\frac{f_{D_s^*}}{2\sqrt{3}}, \eqno{(6)} $$ where $x$, $y$ denote the momentum fractions of $c$, $s$ quarks in the $B_c$ and $D_s^*$ mesons, $f_{B_c}$ and $f_{D_s^*}$ are decay constants of $B_c$ and $D_s^*$ respectively, $$ \epsilon_{B_c}=\frac{m_c}{m_{B_c}}, \eqno{(7)} $$ $$ \epsilon_{D_s^*}=\frac{m_{D_s^*}-m_c}{m_{D_s^*}}. \eqno{(8)} $$ The spinor part of $B_c$ and $D_s^{*}$ [14] are $$ \frac{(\not p+m_{B_c})\gamma_5}{\sqrt{2}}, \eqno{(9)} $$ $$ \frac{(\not q-m_{D_s^*})\not \epsilon}{\sqrt{2}}, \eqno{(10)} $$ which come from the matrix structures of $B_c$ and $D_s^*$ meson wave functions, while $p$ and $q$ are the momenta of the $B_c$ and $D_s^*$ respectively, and $\epsilon$ is the polarization vector of $D_s^{*}$. \begin{flushleft} \subsection*{II ( i ). Electromagnetic penguin contribution} \end{flushleft} The relevant Feynman diagrams which contribute to the short distance electromagnetic penguin process $b\rightarrow s\gamma$ are illustrated as the blob of Fig.1. In the evaluation, we at first integrate out the top quark and the weak $W$ bosons at $\mu=m_W$ scale, generating an effective five-quark theory. By using the renormalization group equation, we run the effective field theory down to b-quark scale to give the leading log QCD corrections. After applying the full QCD equations of motion [15], a complete set of dimension-6 operators relevant for $b\rightarrow s\gamma$ decay can be chosen as $O_1$ - $O_8$, which have been given in the Refs.[1, 9]. The effective Hamiltonian appears at the $W$ scale is given as $$ H_{eff}=\frac{4G_F}{\sqrt{2}}V_{tb}V_{ts}^* \sum\limits_{i=1}\limits^{8}C_i (m_W)O_i(m_W). \eqno{(11)} $$ The coefficients of 8 operators are: $$ C_i(m_W) = 0, i= 1, 3, 4, 5, 6, C_2(m_W) = -1, \eqno{(12)} $$ $$ C_7(m_W) =\frac{1}{2}A(x)-(\frac{f}{f'})^2[B(y)-\frac{1}{6}A(y)], \eqno{(13)} $$ $$ C_8(m_W)=\frac{1}{2}D(x)+(\frac{f}{f'})^2[\frac{1}{6}D(y)-E(y)], \eqno{(14)} $$ where functions $A$, $B$, $D$ and $E$ are defined in the Ref.[1], $x=(\frac{m_t}{m_W})^2$, $y=(\frac{m_t}{m_{\pi_p}})^2$. The running of the coefficients of operators from $\mu=m_W$ to $\mu= m_b$ was well described in Ref.[16]. After renormalization group running we have the QCD corrected coefficients of operators at $\mu= m_b$ scale. $$ C_7^{eff}(m_b) =\varrho^{-\frac{16}{23}}C_7(m_W)+\frac{8}{3} (\varrho^{-\frac{14}{23}}-\varrho^{-\frac{16}{23}})C_8(m_W)+C_2(m_W)\sum \limits_{i=1}\limits^{8}h_i \varrho^{-a_i}, \eqno{(15)} $$ with $$ \varrho = \frac{\alpha_s(m_b)}{\alpha_s(m_W)}, \eqno{(16)} $$ $$ h_i=(\frac{626126}{272277}, -\frac{56281}{51730}, -\frac{3}{7}, -\frac{1}{14} , -0.6494, -0.0380, -0.0186, -0.0057), \eqno{(17)} $$ $$ a_i=(\frac{14}{23}, \frac{16}{23}, \frac{6}{23}, -\frac{12}{23}, 0.4086, -0.4230, -0.8994, 0.1456). \eqno{(18)} $$ Now we write down the amplitude of Fig.1 as $$ \begin{array}{ll} M_a=&\int^1_0 dx_1 dy_1 \Phi_{D_s^*}(y_1)\Phi_{B_c}(x_1)\frac{-iG_F} {\sqrt{2}}V_{tb} V_{ ts}^*C_7^{eff}(m_b)m_be\frac{\alpha_s(m_b)}{2\pi}C_F \\ &\{T_r[(\not q-m_{D_s}^{*})\not\epsilon\sigma_{\mu\nu}(1+\gamma_5) k^{\nu}\eta^{\mu}(\not p-y_1\not q+m_b)\gamma_{\alpha}(\not p+m_{B_c}) \gamma_5\gamma^{\alpha}]\frac{1}{D_1D_3} \\ &+Tr[(\not q-m_{D_s^{*}})\not\epsilon\gamma_{\alpha}(\not q-x_1\not p) \sigma_{\mu\nu}(1+\gamma_5)k^{\nu}\eta^{\mu}(\not p+m_{B_c}) \gamma_5\gamma^{\alpha}]\frac{1}{D_2D_3}\}, \end{array} \eqno{(19)} $$ where $\eta$ is the polarization vector of photon, $x_1$, $y_1$ are the momentum fractions shared by charms in $B_c$ and $D_s^*$ respectively, and the color factor $C_F$ = $\frac{4}{3}$. The factors $D_1$, $D_2$ and $D_3$ in eq.(19) are the forms of $$ \begin{array}{l} D_1=(1-y_1)(m_{ B_c}^{ 2}-m_{D_s^*}^2y_1)-m_{ b}^{ 2},{\hskip 8.5cm} (20)\\ D_2=(1-x_1)(m_{ D_s^*}^2-m_{B_c}^2 x_1),{\hskip 9.5cm}(21) \\ D_3=(x_1-y_1)(x_1m_{B_c}^{ 2}-y_1m_{D_s^*}^{ 2}).{\hskip 9cm}(22) \end{array} $$ Now the amplitude $M_a$ can be written as $$ M_a=i\varepsilon_{\mu\nu\alpha\beta}\eta^{\mu}k^{\nu}\epsilon^{\alpha} p^{\beta}f_{ 1}^{ peng} +\eta^{\mu}[\epsilon_{\mu}(m_{B_c}^2-m_{D_s}^2) -(p+q)_{\mu}(\epsilon\cdot k)]f_2^{ peng}, \eqno{(23)} $$ with form factors $$ \begin{array}{ll} f_1^{peng}&=2f_2^{peng}=C\int_{0}^{1}dx_1dy_1\delta(x_1-\epsilon_{B_c}) \delta(y_1-\epsilon_{D_s^*}) \\ &\{[m_{B_c}(1-y_1)(m_{B_c}-2m_{D_s^*}) -m_b(2m_{B_c}-m_{D_s^*})] {\frac{1}{D_1D_3}} - m_{B_c}m_{D_s^*}(1-x_1) {\frac{1}{D_2D_3}}\}, \end{array} \eqno{(24)} $$ where $$ C=\frac{em_bf_{B_c}f_{D_s^*}C_7^{eff}(m_b)C_F\alpha_s(m_b)G_FV_{tb} V_{ts}^*}{12\pi\sqrt{2}}. \eqno{(25)} $$ \begin{flushleft} \subsection*{II ( ii ). The weak annihilation contribution} \end{flushleft} As mentioned in Sec.I, $B_c$ meson is also the unique probe of the weak annihilation mechanism. The leading log QCD-corrected effective weak Hamiltonian of $W$ annihilation is $$ H^{(W)}_{eff}= \frac{G_F}{2\sqrt{2}}V_{cb}V_{cs}^*( c_+O_+ + c_-O_- ) + H.c, \eqno{(26)} $$ with $O_{\pm}=(\overline sb)(\overline cc)\pm(\overline sc)(\overline cb)$, where $(\overline q_1q_2) \equiv \overline q_1\gamma_{\mu}(1-\gamma_5)q_2$, $c_{\pm}$ are Wilson coefficient functions. Using the method developed by H. Y. Cheng $et \ \ al.$ [17], we can get the amplitude of $W$ annihilation diagrams ( see Fig.2 ): $$ M_{ b}^{(W)}=i\varepsilon_{\mu\nu\alpha\beta}\eta^{\mu}k^{\nu}\epsilon^ {\alpha} p^{\beta}f_{1(W)}^{anni}+\eta^{\mu}[\epsilon_{\mu}(m_{B_c}^2-m_{D_s}^2)- (p+q)_{\mu}(\epsilon\cdot k)]f_{2(W)}^{anni}, \eqno{(27)} $$ with $$ f_{1(W)}^{anni}=2\zeta[(\frac{e_s}{m_s}+\frac{e_c}{m_c}) \frac{m_{D_s^*}}{m_{B_c}}+(\frac{e_c}{m_c}+\frac{e_b}{m_b})] \frac{m_{D_s^*}m_{B_c}}{m_{B_c}^2-m_{D_s^*}^2}, \eqno{(28)} $$ $$ f_{2(W)}^{anni}=-\zeta[(\frac{e_s}{m_s}-\frac{e_c}{m_c})\frac{m_{D_s^*}} {m_{B_c}} +(\frac{e_c}{m_c}-\frac{e_b}{m_b})]\frac{m_{D_s^*}m_{B_c}} {m_{B_c}^2-m_{D_s^*}^2}, \eqno{(29)} $$ where $$ \zeta=ea_2\frac{G_F}{\sqrt{2}}V_{cb}V_{cs}^*f_{B_c}f_{D_s^*}, \eqno{(30)} $$ and $a_2=\frac{1}{2}(c_--c_+)$ is a calculable coefficient in the nonleptonic $B$ decays. Using the Feynman rules given in eq. ( 1 ) and eq. ( 2 ), the leading log QCD-corrected effective weak Hamiltonian of $\pi_p^{\pm}$ annihilation is given as $$ H^{(\pi_p)}_{eff}=-V_{cb}V_{cs}^*(\frac{f}{vf'})^2 \frac{1}{2m^2_{\pi_P}}( c_+ O_+ + c_-O_-) + H.c. \eqno{(31)} $$ Using the same method as the above, we can write down the amplitude of $\pi_p^{\pm}$ annihilation diagrams ( see Fig.2 ) as $$ M_b^{(\pi_p)}=i\varepsilon_{\mu\nu\alpha\beta}\eta^{\mu}k^{\nu}\epsilon^ {\alpha}p^{\beta}f_{1(\pi_p)}^{anni} + \eta^{\mu}[\epsilon_{\mu}(m_{B_c}^2 -m_{D_s^*}^2)-(p+q)_{\mu}(\epsilon \cdot k)]f_{ 2(\pi_p)}^{ anni}, \eqno{(32)} $$ with $$ f_{1(\pi_p)}^{anni}=2\zeta^{'}[(\frac{e_s}{m_s}+\frac{e_c}{m_c}) \frac{m_s-m_c}{m_{B_c}}+(\frac{e_b}{m_b}+\frac{e_c}{m_c}) \frac{m_b-m_c}{m_{B_c}}]\frac{m_{B_c}m_{D_s^*}}{m_{B_c}^2- m_{D_s^*}^2}, \eqno{(33)} $$ $$ f_{2(\pi_p)}^{ anni}=\zeta^{'}[(\frac{e_s}{m_s}+\frac{e_c}{m_c}) \frac{m_{D_s^*}}{m_{B_c}}+(\frac{e_b}{m_b}+\frac{e_c}{m_c}) ]\frac{m_{D_s^{*}}m_{B_c}}{m_{ B_c}^{ 2} -m_{ D_s^{*}}^{ 2}}, \eqno{(34)} $$ where $$ \zeta^{'}=ea_2V_{cb}V_{cs}^*(\frac{f}{vf'})^2\frac{1}{4m_{\pi_p}^2}f_{B_c} f_{D_s^*}(m_{B_c}^2+m_{D_s^*}^2). \eqno{(35)} $$ The total annihilation amplitude ( Fig.2 ) is the form of $$ \begin{array}{lll} M_b&=&M_b^{(W)}+M_b^{(\pi_p)} \\ &=&i\varepsilon_{\mu\nu\alpha\beta}\eta^{\mu}k^{\nu}\epsilon^{\alpha} p^{\beta}f_1^{anni}+\eta^{\mu}[\epsilon_{\mu}(m_{B_c}^2-m_{D_s}^2) -(p+q)_{\mu}(\epsilon \cdot k)]f_2^{anni}, \end{array} \eqno{(36)} $$ where $$ f_1^{anni} = f_{1(W)}^{anni} +f_{1(\pi_p)}^{anni}, \eqno{(37)} $$ $$ f_2^{anni} = f_{2(W)}^{anni} +f_{2(\pi_p)}^{anni}. \eqno{(38)} $$ Finally, we estimate another possible long-distance effect, namely the vector -meson-dominance ( VMD ) contribution which was advocated by Golowich and Pakvasa [18]. VMD implies that a possible contribution to $B_c\rightarrow D_s^*\gamma$ comes from the $B_c\rightarrow D_s^*J/\psi(\psi')$ followed by $J/\psi(\psi')\rightarrow\gamma$ conversion. As discussed in Refs.[19, 20], using factorization approach, the VMD amplitude ( Fig.3 ) is $$ M^{VMD}=i\varepsilon_{\mu\nu\alpha\beta}\eta^{\mu}k^{\nu}\epsilon^ {\alpha} p^{\beta}f_1^{VMD}+\eta^{\mu}[\epsilon_{\mu}(m_{B_c}^2-m_{D_s}^2)- (p+q)_{\mu}(\epsilon\cdot k)]f_2^{VMD}, \eqno{(39)} $$ with $$ \begin{array}{ll} f_1^{VMD}&=eG_FV_{cb}V_{cs}^* \{\sqrt{2}a_2\frac{1}{m_{B_c}+m_{D_s^*}} (\frac{f_{J/\psi}m_{J/\psi}}{g_{\gamma J/\psi}}+\frac{f_{\psi'}m_{\psi'}} {g_{\gamma\psi'}})\\ & + a_1f_{D_s^*}m_{D_s^*}[\frac{1}{(m_{B_c}+m_{J/\psi}) g_{\gamma J/\psi}}+\frac{1}{(m_{B_c}+m_{\psi'})g_{\gamma\psi'}}]\} V^{BD}(0), \end{array} \eqno{(40)} $$ $$ f_2^{VMD}=-\frac{1}{2}\frac{A_2^{BD}(0)}{V^{BD}(0)}f_1^{VMD}, \eqno{(41)} $$ where $a_1=\frac{1}{2}(c_++c_-)$, and $V^{BD}(0)$ and $A_2^{BD}(0)$ are form factors. \begin{flushleft} \section*{III. Numerical results} \end{flushleft} We will use the following values for various quantities to carry on our calculations. ( i ). Decay constants for mesons. Here we use $$ f_{D_s^*}=f_{D_s}=344 MeV [21], \ \ f_{B_c}=500 MeV [22], \ \ f_{J/\psi}=395 MeV [20], \ \ f_{\psi'}=293 MeV [20]. $$ ( ii ). Meson mass and the constituent quark mass [23, 24] $$ m_{B_c}=6.27 GeV, \ \ m_{D_s^*}=2.11 GeV, \ \ m_b=4.7 GeV, \ \ m_c=1.6 GeV, $$ $$ m_s=0.51 GeV, \ \ m_{J/\psi}=3.079 GeV, \ \ m_{\psi'}=3.685 GeV. $$ We also use $m_{B_c}\approx m_b+m_c$, $m_{D_s^*}\approx m_s+m_c$ in our calculations. ( iii ). $a_1$ and $a_2$ have been estimated very recently in Ref.[17] according to the CLEO data [25] on $B\rightarrow D^* \pi(\rho)$ and $B\rightarrow J/\psi K^*$. Here we take $$ a_1=1.01, \ \ a_2=0.21. $$ ( iv ). CKM matrix elements [24]. Here we use $$ \vert V_{cb}\vert=0.04, \ \ \vert V_{ts}\vert=\vert V_{cb}\vert, \ \ \vert V_{cs}\vert=0.9745, \ \ \vert V_{tb}\vert=0.9991. $$ ( v ). The QCD coupling constant $\alpha_s(\mu)$ at any renormalization scale, can be calculated from $\alpha_s(m_Z)=0.117$ via $$ \alpha_s(\mu)=\frac{\alpha_s(m_Z)}{1-(11-\frac{2}{3}n_f)\frac{\alpha_s(m_Z)} {2\pi}\ln(\frac{m_Z}{\mu})}. $$ We obtain $$ \alpha_s (m_b)=0.203, \ \ \ \alpha_s(m_W)=0.119. $$ ( vi ). The Ref.[10] gives a constraint on $m_{\pi_p}$ in the allowed parameter space of the model: $\frac{1}{2}m_Z<m_{\pi_p}<1 TeV$. Here we take $$ m_{\pi_p}=( 50\sim 1000 ) GeV. $$ ( vii ). With the constraint of $f^2+f^{'2}=v^2$ and the chiral perturbation theory in Ref.[10], we can get $0.115\leq \frac{f}{f'} \leq 1.74$. Here we take $$ \frac{f}{f'}=0.115 $$ in our calculations. ( viii ). The form factors $V(0)$ and $A_2(0)$ appearing in the two-body decays of $B$. From Ref.[26], we take $$ V^{BD}(0)=0.30, \ \ A_2^{BD}(0)=0.20. $$ We present the form factors $f_i$ ( $f_1^{peng}$, $f_2^{peng}$, $f_1^{anni}$, $f_2^{anni}$ ) in the SM and TCMLSM in Table 1, so do the decay widths in Table 2 using the amplitude formula $$ \Gamma(B_c\rightarrow D_s^*\gamma) = \frac{(m_{B_c}^2-m_{D_s^*}^2)^3} {32\pi m_{B_c}^3}(f_1^2 +4f_2^2). $$ The calculated results indicate that the VMD effects are large near the pole and which can not be neglected. The calculated results are $$ f_1^{VMD}=6.73\times 10^{-10}, \ \ f_2^{VMD}=-2.24\times 10^{-10}, $$ $$ \Gamma^{VMD}=1.12\times 10^{-18} GeV. $$ The lifetime of $B_c$ is given in Ref.[27]. In this paper we use $$ \tau_{B_c}=( 0.4 ps\sim 1.35 ps ) $$ to estimate the branching ratio BR ( $B_c\rightarrow D_s^*\gamma$ ) which is a function of $\tau_{B_c}$. The results are given in Table 3. \begin{flushleft} \section*{IV. Discussion } \end{flushleft} Applying PQCD, we have studied two mechanisms which contribute to the process $B_c\rightarrow D_s^*\gamma$. For the short-distance one ( Fig.1 ) induced by electromagnetic penguin diagrams, the momentum square of the hard scattering exchanged by gluon is about $3.6 GeV^2$ which is large enough for PQCD analyzing. The hard scattering process can not be included conveniently in the soft hadronic process described by the wave-function of the final bound state, which is one important reason that we can not apply the commonly used models with spectator [28] to the two body $B_c$ decays. There is no phase-space for the propagators appearing in Fig.1 to go on-shell, consequently, unlike the situation in the Ref.[6], the imaginary part of $M_a$ is absent. Another competitive mechanism is the weak annihilation. In the SM, we find this mechanism is as important as the former one ( they can contribute with the same order of magnitude ). This situation is distinct from that of the radiative weak $B^{\pm}$ decays which is overwhelmingly dominated by electromagnetic penguin. This is due to two reasons: one is that the compact size of $B_c$ meson enhances the importance of annihilation decays; the other comes from the Cabibbo allowance: in $ B_c \rightarrow D_s^* \gamma$, the CKM amplitude of weak annihilation is $\vert V_{cb}V_{cs}^*\vert $, but in $ B^{\pm} \rightarrow K^{\pm}\gamma$, the CKM part is $ \vert V_{ub}V_{us}^* \vert $ which is much smaller than $\vert V_{cb}V_{cs}^*\vert$. Particularly, the VMD contribution is found not small. This situation is quite different from the cases $B\rightarrow J/\psi K(K^*)$ and $B\rightarrow J/\psi\rho$ [19, 20]. The reason comes from that although the coupling of $J/\psi(\psi') - \gamma$ is small ( $e/g_{\gamma J/\psi(\psi')}\approx 0.025(0.016)$ ), the $J/\psi(\psi')$ resonance effect can be very large. In addition, we find that the modification of $B_c\rightarrow D_s^* \gamma $ from $\pi _p$ in the TCMLSM is small for the allowed range of mass of $\pi_p$ ( with $\frac{f}{f'}$ fixed ). This situation is quite different from that of Ref.[9], in which the size of contribution to the rare decay of $b\rightarrow s\gamma$ from the PGBs strongly depends on the values of the masses of the charged PGBs. The difference is mainly due to the small value of $\frac{f}{f'}$ which leads to the small modification from $\pi_p$ in the TCMLSM. However, in the OGTM, such suppression factor $\frac{f}{f'}$ does not exist. In our calculations, we take $\frac{f}{f'}$=0.115 as the input parameter. When $\frac{f}{f'}$ is taken properly larger ( without exceeding the constraint: $0.115\leq \frac{f}{f'} \leq 1.74$ ), the calculated results remain unchanged basically. In view of the above situation, it seems to indicate that the window of process $B_c\rightarrow D_s^*\gamma$ is close for the TCMLSM. But in our calculations, besides the peaking approximation of the meson wave functions, the theoretical uncertainties are neglected, such as that of $\alpha ( m_Z ) $, next-to-leading log QCD contribution [29], QCD correction from $m_t$ to $m_W$ [30], etc. When the more reliable estimation is available within the next few years, one can, in principle, make the final decision whether the window for TCMLSM is open or close. \vspace{1cm} \noindent {\bf ACKNOWLEDGMENT} This work was supported in part by the National Natural Science Foundation of China, and by the funds from Henan Science and Technology Committee. \vspace{1cm} \begin {center} {\bf Reference} \end {center} \begin{enumerate} \item B. Grinstein $et \ \ al.$, Nucl. Phys. B 339 ( 1990 ) 269. \item D. S. Du and Z. Wang, Phys. Rev. D 39 ( 1989 ) 1342; K. Cheng, T. C. Yuan, Phys. Lett. B 325 ( 1994 ) 481, Phys. Rev. D 48 ( 1994 ) 5049; G.R. Lu, Y.D. Yang and H.B. Li, Phys. Lett. B341 (1995)391, Phys. Rev. D51 (1995)2201. \item J. Tang, J. H. Liu and K. T. Chao, Phys. Rev. D51 (1995)3501; K. C. Bowler $ et \ \ al.$, Phys. Rev. Lett. 72 ( 1994 ) 1398. \item H. Leutwyler and M. Roos, Z. Phys. C 25 ( 1984 ) 91. \item M. Neubert, Phys. Rep. 245 ( 1994 ) 1398. \item A. Szczepaiak $et \ \ al.$, Phys. Lett. B 243 ( 1990 ) 287; C. E. Carlson and J. Milana, Phys. Lett. B 301 ( 1993 ) 237, Phys. Rev. D 49 ( 1994 ) 5908, $ibid$ 51 ( 95 ) 450; H-n. Li and H. L. Yu, Phys. Rev. Lett. 74 ( 1995 ) 4388. \item D. S. Du, $et \ \ al.$, BIHEP-TH-95-38 ( hep-ph/9603291 ) \item S. J. Brodsky and G. P. Lepage, Phys. Rev. D 22 ( 1980 ) 2157. \item Cai-Dian L\"u and Zhenjun Xiao, Phys. Rev. D 53 ( 1996 ) 2529. \item C. D. Carone and H. Georgi, Phys. Rev. D 49 ( 1994 ) 1427. \item C. D. Carone and E. H. Simmons, Nucl. Phys. B 397 ( 1993 ) 591; E. H. Simmons, Nucl. Phys. B 312 ( 1989 ) 253; S. Samuel, $ibid$, B 347 ( 1990 ) 625; A. Kagan and S. Samuel, Phys. Lett. B 252 ( 1990 ) 605, B 270 ( 1991 ) 37. \item S. J. Brodsky and C. R. Ji, Phys. Rev. Lett. 55 ( 1985 ) 2257. \item Hsiang-nan Li, Phys. Rev. D 52, ( 1995 ) 3958. \item C. E. Carlson, J. Milana, Phys. Rev. D 51, ( 1995 ) 4950. \item H. D. Politzer, Nucl. Phys. B 172 ( 1980 ) 349; H. Simma, Preprint, DESY 93-083. \item M. Misiak, Phys. Lett. B 269 ( 1991 ) 161; K. Adel, Y. P. Yao, Mod. Phys. Lett. A 8 ( 1993 ) 1679; Phys. Rev. D 49 ( 1994 ) 4945; M. Ciuchini $ et \ \ al.$, Phys. Lett. B 316 ( 1993 ) 127. \item H. Y. Cheng $et \ \ al.$, Phys. Rev. D 51 ( 1995 ) 1199. \item E. Golowich and S. Pakvasa, Phys. Lett. B 205 ( 1988 ) 393. \item M. Bauer, B. Stech and M. Wirbel, Z. Phys.C 34 ( 1987 ) 103; E. Golowich and S. Pakvasa, Phys. Rev. D 51 ( 1995 ) 1215. \item H. Y. Cheng, Phys. Rev. D 51 ( 1995 ) 6228. \item A. Aoki $et \ \ al.$, Prog. Theor. Phys. 89 ( 1993 ) 137; D. Acosta $et \ \ al.$, CLNS 93/1238; J. Z. Bai $ et \ \ al.$, BES Collaboration Phys. Rev. Lett. 74 ( 1995 ) 4599. \item W. Buchm\"uiller and S-H. HTye, Phys. Rev. D 24 ( 1994 ) 132; A. Martin, Phys. Lett. B 93 ( 1980 ) 338; C. Quigg and J. L. Rosner, Phys. Lett. B 71 ( 1977 ) 153; E. Eicheten $et \ \ al.$, Phys. Rev. D 17 ( 1978 ) 3090. \item W. Kwong and J. L. Rosner, Phys. Rev. D 44 ( 1991 ) 212. \item Particle Data Group, L. Montanet $et \ \ al.$, Phys. Rev. D 50 ( 1994 ) 1173. \item CLEO Collaboration, M. S. Alam $et \ \ al.$, Phys. Rev. D 50 ( 1994 ) 43. \item M. Neubert, CERN-TH/96-55 (hep-ph/9604412). \item C. Quigg, FERMILAB-Conf-93/265-T; C. H. Chang and Y. Q. Chen, Phys. Rev. D49 ( 1994 ) 3399; P. Colangelo $et \ \ al.$, Z. Phys. C 57 ( 1993 ) 43. \item M. Bauer $et \ \ al.$, Z. Phys. C29 ( 1985 ) 637; B. Grinstein $et \ \ al.$, Phys. Rev. D39 ( 1987 ) 799. \item A. J. Buras $et \ \ al.$, Nucl. Phys. B 370 ( 1992 ) 69; Addendum, $ibid$, B 375 ( 1992 ) 501, B 400 ( 1993 ) 37 and B 400 ( 1993 ) 75; M. Ciuchini $et \ \ al.$, Phys. Lett. B 301 ( 1993 ) 263, Nucl. Phys. B 415 ( 1994 ) 403. \item C. S. Gao, J. L. Hu, C. D. L\"u and Z. M. Qiu, Phys. Rev. D52 (1995)3978. \end{enumerate} \newpage \begin{table}[h] \caption{Form factors in the SM and TCMLSM. $f^{peng}$ and $f^{anni}$ represent form factors through electromagnetic penguin process and through weak annihilation process respectively.} \begin{center} \begin{tabular}{|c|c|c|} \hline $f_i$ & SM & TCMLSM \\ \hline $f_1^{peng}$ & $-3.05\times 10^{-10}$ & $(-3.02\sim -3.05)\times 10^{-10}$ \\ \hline $f_2^{peng}$ & $-1.52\times 10^{-10}$ &$ -1.52\times 10^{-10}$ \\ \hline $f_1^{anni}$ & $7.10\times 10^{-10}$ & $ 7.10\times10^{-10}$ \\ \hline $f_2^{anni}$ & $-1.70\times 10^{-10}$ & $ -1.70\times 10^{-10}$ \\ \hline \end{tabular} \end{center} \end {table} \begin{table}[h] \caption{The decay rates in the SM and TCMLSM. The $\Gamma^{peng}$, $\Gamma^ {anni}$ and $\Gamma^{total}$ represent $\Gamma$ ( $B_c\rightarrow D_s^* \gamma$ ) through electromagnetic penguin process, through weak annihilation process and penguin + annihilation respectively.} \begin{center} \begin{tabular}{|c|c|c|} \hline $\Gamma(B_c \rightarrow D_s^* \gamma)$ & SM & TCMLSM \\ \hline $\Gamma^{peng}$(GeV) &$ 3.18 \times 10^{-19}$ & $ ( 3.12\sim 3.17 ) \times 10^{-19}$ \\ \hline $\Gamma^{anni}$(GeV) & $1.06\times 10^{-18}$ & $ 1.06 \times 10^{-18}$ \\ \hline $\Gamma^{total}$(GeV) & $4.03\times 10^{-18}$ & $ 4.03\times 10^{-18}$ \\ \hline \end{tabular} \end{center} \end{table} \begin{table}[h] \caption{The branching ratio ( $B_c\rightarrow D_s^*\gamma$ ). The $BR^{SM}_ {total}$ and $BR^{TCMLSM}_{total}$ represent the branching ratio ( $B_c \rightarrow D_s^*\gamma$ ) in the SM and TCMLSM respectively.} \begin{center} \begin{tabular}{|c|c|c|c|} \hline $\tau_{B_c} $ & 0.4ps & 1.0ps & 1.35ps \\ \hline $ BR^{SM}_{total} $ & $2.44\times 10^{-6}$ & $ 6.12\times 10^{-6}$ & $8.27\times 10^{-6}$ \\ \hline $BR^{TCMLSM}_{total}$ &$2.44\times 10^{-6}$ &$6.12\times 10^{-6}$ &$8.27\times 10^{-6}$ \\ \hline \end{tabular} \end{center} \end{table} \vspace{1cm} \begin{center} {\bf Figure captions} \end{center} Fig.1: The Feynman diagrams which contribute to the rare radiative decay $B_c\rightarrow D_s^*\gamma$ through electromagnetic penguin process. The blob represents the electromagnetic penguin operators contributing to $b\rightarrow s\gamma$, $x_2 p$ and $x_1 p$ are momenta of b and c quarks in the $B_c$ meson respectively, $y_2 q$ and $y_1q$ are momenta of s and c quarks in the $D_s^*$ meson, respectively. Fig.2: The Feynman diagrams which contribute to the rare radiative decay $B_c\rightarrow D_s^*\gamma$ through weak annihilation process. In the SM, there is only $W^{\pm}$ annihilation, in the TCMLSM, there are both $W^{\pm}$ and $\pi_p^{\pm}$ annihilations. Fig.3: VMD processes which contribute to $B_c\rightarrow D_s^*\gamma$ with the vector-meson intermediate states $J/\psi(\psi')$. \newpage \begin{picture}(30,0) \setlength{\unitlength}{0.1in} \put(5,-15){\line(1,0){20}} \put(5,-20){\line(1,0){20}} \put(3,-18){$B_c$} \put(26,-18){$D_s^*$} \multiput(10,-15.8)(0,-0.8){6}{$\varsigma$} \multiput(14,-15)(1,1){5}{\line(0,1){1}} \multiput(13,-15)(1,1){6}{\line(1,0){1}} \put(14,-15){\circle*{3}} \put(11,-18){$q_G$} \put(16.2,-14){( $W^{\pm}$, $\pi_p^{\pm}$ , c, t )} \put(5,-16.5){$x_2 p$} \put(5,-21.5){$x_1 p$} \put(23,-16.5){$y_2 q$} \put(23,-21.5){$y_1 q$} \put(11,-14.5){$l_1 $} \put(29,-25.2){Fig.1} \put(21,-8){k} \put(19,-10){$\gamma$} \put(35,-15){\line(1,0){20}} \put(35,-20){\line(1,0){20}} \multiput(47,-15.8)(0,-0.8){6}{$\varsigma$} \multiput(44,-15)(1,1){5}{\line(0,1){1}} \multiput(43,-15)(1,1){6}{\line(1,0){1}} \put(48,-18){$q_G$} \put(35,-16.5){$x_2 p$} \put(35,-21.5){$x_1 p$} \put(53,-16.5){$y_2 q$} \put(53,-21.5){$y_1 q$} \put(46,-14.5){$l_2 $} \put(44,-15){\circle*{3}} \put(51,-8){k} \put(49,-10){$\gamma$} \put(5,-35){\line(5,-4){5}} \put(5,-43){\line(5,4){5.5}} \multiput(10.4,-39.5)(1,0){7}{V} \put(17.5,-38.5){\line(5,-4){5.5}} \put(18,-39.1){\line(5,4){5}} \multiput(51.7,-41)(1,1){3}{\line(0,1){1}} \multiput(50.7,-41)(1,1){4}{\line(1,0){1}} \put(3,-39){$B_c$} \put(24,-39){$D_s^*$} \put(11,-42){ $W^{\pm}$, $\pi_p^{\pm}$ } \put(11,-33){$\gamma$} \put(35,-35){\line(5,-4){5}} \put(35,-63){\line(5,4){5.5}} \multiput(40.4,-59.5)(1,0){7}{V} \put(47.5,-58.5){\line(5,-4){5.5}} \put(48,-59.1){\line(5,4){5}} \multiput(8.5,-61)(1,-1){3}{\line(0,-1){1}} \multiput(7.5,-61)(1,-1){4}{\line(1,0){1}} \put(5,-55){\line(5,-4){5}} \put(5,-63){\line(5,4){5.5}} \multiput(10.4,-59.5)(1,0){7}{V} \put(17.5,-58.5){\line(5,-4){5.5}} \put(18,-59.1){\line(5,4){5}} \multiput(51.7,-55.5)(-1,1){4}{\line(0,-1){1}} \multiput(50.7,-55.5)(-1,1){4}{\line(1,0){1}} \put(35,-55){\line(5,-4){5}} \put(35,-43){\line(5,4){5.5}} \multiput(40.4,-39.5)(1,0){7}{V} \put(47.5,-38.5){\line(5,-4){5.5}} \put(48,-39.1){\line(5,4){5}} \multiput(8.5,-37)(1,1){3}{\line(0,1){1}} \multiput(7.5,-37)(1,1){4}{\line(1,0){1}} \put(29,-70){Fig.2} \end{picture} \newpage \begin{picture}(30,0) \setlength{\unitlength}{0.1in} \put(5,-15){\line(1,0){9.5}} \put(14,-16.4){\oval(8,3)[r]} \multiput(9.6,-19)(-0.2,0.5){8}{$v$} \put(6.5,-18){$W^-$} \put(14,-20){\oval(8,4.1)[l]} \put(5,-25){\line(1,0){18}} \put(14,-22.1){\line (1,0){8.5}} \put(18,-18.5){$J/{\psi}(\psi^{'})$} \multiput(19,-16.4)(1,1){3}{\line(0,1){1}} \multiput(18,-16.4)(1,1){4}{\line(1,0){1}} \put(22,-13){$\gamma$} \put(5.5,-14.5){b} \put(13,-14.5){c} \put(13,-21.5){s} \put(13,-17.5){$\overline {c}$} \put(5.5,-24.5){$\overline {c}$} \put(3,-20){$B_c$} \put(23,-24){$D_s^*$} \put(29,-30){$Fig.3$} \multiput(40,-22.1)(0.5,0.3){12}{$v$} \put(41,-20){$W^-$} \put(50,-18){\oval(8,3)[l]} \put(35,-25.1){\line(1,0){14}} \put(35,-22.1){\line (1,0){14}} \put(49,-23.5){\oval(8,3)[r]} \put(54,-24.5){$J/{\psi}(\psi^{'})$} \multiput(54,-24)(1,1){3}{\line(0,1){1}} \multiput(53,-24)(1,1){4}{\line(1,0){1}} \put(57,-20){$\gamma$} \put(35.5,-21.5){b} \put(48,-21.5){c} \put(48,-16){s} \put(48,-19.5){$\overline {c}$} \put(35.5,-24.5){$\overline {c}$} \put(33,-24){$B_c$} \put(51,-18.5){$D_s^*$} \end{picture} \end{document} 
proofpile-arXiv_065-447
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Figure captions} {\bf Fig. 1} Diagrams with four quarks in the final state containing a cut triangular quark loop.\\[2ex] {\bf Fig. 2} Contributions to the asymmetry fragmentation function $F_A(x,Q^2)$ \eref{eq:24} at $Q=M_Z$ using the fragmentation density set of \cite{Bin95}. Solid line: LO. Dashed line: NLO. Dotted line: NNLO. The exprimental data are taken from OPAL \cite{Ake95}.\\[2ex] {\bf Fig. 3} The ratio $R_A(x,Q^2)$ \eref{eq:25} at $Q=M_Z$ using the fragmentation density set of \cite{Bin95}. Short dashed line: LO. Solid line: NLO. Dotted line: NNLO. The exprimental data are taken from OPAL \cite{Ake95}.
proofpile-arXiv_065-448
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} Galaxies are open systems and the properties of the interstellar gas are regulated by the internal sources of energy and interactions with other galaxies. There are a series of different mechanisms, for both isolated and interacting galaxies, that are able to accumulate large gas masses and create star-forming clouds in relatively short time scales. These include cloud collisions, gravitational and thermal instabilities, Parker instabilities, gas flows in a bar potential, tidal interactions, direct galaxy collisions, and mergers (some of these mechanisms are discussed in this volume by Boeker {\it et al.}, Borne {\it et al.}, Dultzin-Hacyan, Elmegreen, Friedli \& Martinet, Lamb {\it et al.}, and Moss \& Whittle). Any one, or a combination, of these processes could be operative in different locations and at different moments in the host galaxy, and star formation is the end product of a series of successive condensations of the interstellar medium. Once a cloud is formed, however, a distinction should be made between molecular and self-gravitating clouds (see Franco \& Cox 1986). The criterion to form molecular clouds is high opacity in the UV, to prevent molecule photo-destruction, and this is achieved at column densities above $N_{\tau} \sim 5\times 10^{20} (Z/Z_{\odot})^{-1}$ cm$^{-2}$, where $Z$ is the metallicity and $Z_{\odot}$ is the solar value. In contrast, self-gravity becomes dominant when the column density becomes larger than $N_{sg} \sim 5\times 10^{20} (P/P_{\odot})^{1/2}$ cm$^{-2}$, where $P$ is the interstellar pressure and $P_{\odot}$ is the value at the solar circle. These two values are similar at the location of the Sun, but $N_{\tau} < N_{sg}$ in the inner Galaxy and $N_{\tau} > N_{sg}$ in the outer parts of the Milky Way. This difference has important consequences and may explain the observed radial trends of molecular gas in spirals: it is easier to form molecular clouds in the internal, chemically evolved, parts of spiral galaxies. In any case, the transformation of gas into stars is due to a gravitational collapse and self-gravity defines the structure of the star forming clouds. \section{Self-gravity} The formation of stellar groups (or isolated stars, if any) occurs in the densest regions, the cores, of massive and self-gravitating clouds. In our Galaxy, the $average$ densities for giant molecular cloud complexes is in the range $10^2$ to $10^3$ cm$^{-3}$, but the actual densities in the dense cores is several orders of magnitude above these values: close to about $\sim 10^6$ cm$^{-3}$ ({\it e.g.},\ Bergin {\it et al.}\ 1996; see recent review by Walmsley 1995). Moreover, recent studies of young stellar objects suggest the existence of even denser gas, with values in excess of $10^8$ cm$^{-3}$ (Akeson {\it et al.}\ 1996). Thus, the parental clouds have complex clumpy (and filamentary) structures, with clump-interclump density ratios of about $\sim 10^2$, or more, and temperatures ranging between 10 and $10^2$ K. In addition, the existence of large non-thermal velocities, of several km s$^{-1}$, and strong magnetic fields, ranging from tens of $\mu$G to tens of mG (see Myers \& Goodman 1988 and references therein), indicate large $total$ internal pressures, up to more than five orders of magnitude above the ISM pressure at the solar neighborhood (which is about $10^{-12}$ dyn cm$^{-2}$). A simple estimate for isothermal, spherically symmetric, clouds (with a central core of constant density $\rho_c$ and radius $r_c$, and an external diffuse envelope with a density stratification $\rho = \rho_c (r/r_c)^{-2}$), indicates that self-gravity provides these large total pressure values (see Garc\'{\i}a-Segura \& Franco 1996). In hydrostatic equilibrium, the pressure difference between two positions located at radii $r_1$ and $r_2$ from the center of the core is given by $\Delta P = -\int^{r_2}_{r_1} \rho g_r dr$, where $g_r$ is the gravitational acceleration in the radial direction. The total pressure at the core center is \begin{equation} P(0)=P_0= \frac{2 \pi G}{3} \rho_c^2 r_c^2 + P(r_c)= \frac{8}{5} P(r_c) \simeq 2\times 10^{-7} \ n_6^2 r_{0.1}^2 \ \ \ \ {\rm dyn \ cm^{-2}}, \end{equation} where $G$ is the gravitational constant, $P(r_c)$ is the pressure at the core boundary $r=r_c$, $n_6=n_c/10^6$ cm$^{-3}$, and $r_{0.1}=r_c/0.1$ pc. The corresponding core mass is \begin{equation} M_c\simeq \left( \frac{\pi P_0}{G}\right)^{1/2} r_c^2 \sim 10^2 \ P_7^{1/2} r_{0.1}^2 \ \ \ \ {\rm M_{\odot}}, \end{equation} where $P_7= P_0/10^{-7}$ dyn cm$^{-2}$. For $P_7\sim 1$ and a typical core size for galactic clouds, $r_{0.1}\sim 1$ (see Walmley 1995), gives a value similar to the observationally derived core masses; in the range of 10 to 300 M$_{\odot}$ ({\it e.g.},\ Snell {\it et al.}\ 1993). The pressure inside the core varies less than a factor of two between the center and $r=r_c$. Taking $r_{0.1}=1$ and the maximum core density value, $n_c \sim 5\times 10^6$ cm$^{-3}$ ({\it e.g.},\ Bergin {\it et al.}\ 1996), the upper bound for the expected core pressures is about $P_0 \simeq 5\times 10^{-6}$ dyn cm$^{-2}$. The large range in observed cloud properties obviously results in pressure fluctuations of a few orders of magnitude (both, from cloud to cloud and inside any given cloud), and it is meaningless to define an ``average'' cloud pressure value. Actually, given that star forming clouds have nested structures, in which dense fragments are embedded in more diffuse envelopes, different cloud locations have different total pressures. Also, the expected range of cloud pressures in our Galaxy should probably span from the ISM values, $P_7\sim 10^{-5}$, at the very external cloud layers, up to $P_7\sim 10$ inside the most massive star forming cores. \section{Stellar radiation: HII regions and cloud destruction} The initial structure and pressure of the gas in a star forming cloud is defined by self-gravity. Once young stars appear, the new energy input modifies the structure and evolution of the cloud. Low-mass stars provide a small energy rate and affect only small volumes, but their collective action may provide partial support against the collapse of their parental clouds, and could regulate some aspects of the cloud evolution (Norman \& Silk 1980; Franco \& Cox 1983; Franco 1984; McKee 1989; see also the paper by Bertoldi \& McKee in this volume). In contrast, the strong radiation fields and fast stellar winds from massive stars are able to excite large gas masses and can even disrupt their parental clouds ({\it e.g.},\ Whitworth 1979; Franco {\it et al.}\ 1994). Also, they are probably responsible for both stimulating and shutting off the star formation process at different scales. The combined effects of supernovae, stellar winds, and H II region expansion destroy star-forming clouds and can produce, at some distance and later in time, the conditions for further star formation ({\it e.g.},\ Franco \& Shore 1984; Palous {\it et al.}\ 1995). Thus, the transformation of gas into stars may be a self-limited and self-stimulated process (see reviews by Franco 1991, Ferrini 1992, and Shore \& Ferrini 1994). In the case of the dense star-forming cores, the sizes of either HII regions or wind-driven bubbles are severely reduced by the large ambient pressure (Garc\'{\i}a-Segura \& Franco 1996). In fact, the pressure equilibrium radii of ultra-compact HII regions are actually indistinguishable from those of ultra-compact wind-driven bubbles. When pressure equilibrium is reached, the UCHII radius is \begin{equation} R_{{\rm UCHII,eq}} \approx 2.9 \times 10^{-2} \,\, F_{48}^{1/3} \,\, T_{{\rm HII},4}^{2/3} \,\, P_7^{-2/3} \ \ \,\,{\rm pc} , \label{Rsequnits} \end{equation} where $F_{48}$ is the total number of ionizing photons per unit time in units of $10^{48}$ s$^{-1}$, and $T_{{\rm HII},4} = T / 10^4$ K. For the case of a strong wind evolving in a high-density molecular cloud core, the equilibrium radius of a radiative bubble is \begin{equation} R_{,{\rm WDB,eq}}=\left[ \frac{\dot{M} \,v_{\infty}}{4 \,\pi \,P_0} \right]^{1/2} \simeq 2.3\times 10^{-2} \left[ \frac{\dot{M_6} \,\,v_{\infty,8}}{P_7} \right]^{1/2} \ \ \,\,{\rm pc} , \label{Req} \end{equation} where the mass loss rate is $\dot{M_6}=\dot{M}/10^{-6}$ M$_\odot$ yr$^{-1}$, and the wind velocity is $v_{\infty,8}= v_{\infty}/10^8$ cm s$^{-1}$. Thus, for dense cores with $r_c\sim 0.1$ pc, the resulting UCHIIs and wind-driven bubbles can reach pressure equilibrium without breaking out of the core ({\it i.e.},\ they could be stable and long lived). Recently, Xie {\it et al.}\ (1996) have found evidence indicating that this is probably the case: the smaller UCHII seem to be embedded in the higher pressure cores. If the limit to continued star forming activity inside the core is due to photoionization by these internal H II regions, the maximum number of OB stars is given by the number of H II regions required to completely ionize the core (Franco et al 1994), $N_{OB}=(1-\epsilon)M_{c}/M_{i}$, where $M_{i}$ is the ionized mass. This means that the maximum number of massive stars that can be formed within a core is \begin{equation} N_{OB} \approx 3 {M_{c,2} n_{6}^{3/7} \over F_{48}^{5/7} (c_{s,15} t_{MS,7})^{6/7}} \end{equation} where $M_{c,2}$ is the core mass in $10^{2}$ M$_{\odot}$, $c_{s,15}$ is the HII region sound speed in units of 15 km s$^{-1}$, and $t_{MS,7}$ is the mean OB star main sequence lifetime in $10^{7}$yr. Clearly, for increasing core densities, the value of $R_{0}$ decreases and the resulting number of OB stars increases. In the case of the gas in nuclear regions, due to the intrinsic larger ISM pressures in the inner regions of galaxies, the population of clouds are denser and more compact. The corresponding star forming clouds should also be denser than in the rest of the disk, and a larger number of stars can be formed per unit mass of gas. Thus, {\it nuclear starbursts can be a natural consequence of the higher pressure values} (a bursting star formation mode can also be associated to a delayed energy input, see Parravano 1996). When stars are located near the edge of the core, and depending on the slope of the external density distribution, both HII regions and wind-driven bubbles can accelerate and flare out with a variety of hydro-dynamical phenomena. These include supersonic outflows, internal shocks, receding ionization fronts, fragmentation of the thin shell, etc ({\it e.g.},\ Tenorio-Tagle 1982; Franco {\it et al.}\ 1989, 1990; Garc\'{\i}a-Segura \& Mac Low 1995a,b). Thus, no static solution exists in this case and the pressure difference between the HII regions and the ambient medium begins to evaporate gas from the cloud. This represents a clear and simple physical mechanism for cloud destruction and, as the number of OB stars increases, more expanding H II regions form and limit the rate of new star formation by ionizing the surrounding molecular gas (Franco et al 1994). Eventually, when the whole cloud is completely ionized, star formation ceases. The total cloud mass ionized by an average OB star, integrated over its main sequence lifetime, is \begin{equation} M_{i}(t) \approx \frac{2\pi}{3} R_{0}^{3}\mu_p n_{0} \left\lbrack\left(1+\frac{5c_{s}t_{MS}}{2R_{0}}\right)^{6/5}-1\right\rbrack. \end{equation} where $R_{0}$ is the initial radius at the average cloud density, $n_{0}$, $\mu_p$ is the mass per gas particle, $c_{s}$ is the sound speed in the HII region, and $t_{MS}$ is the main sequence lifetime of the average OB star. For a cloud of mass $M_{GMC}$, with only 10\% of this mass in star-forming dense cores, the number of newly formed OB stars required to completely destroy it is \begin{equation} N_{OB} \sim 30 \frac{M_{GMC,5}n_{3}^{1/5}}{F_{48}^{3/5}(c_{s,15} t_{MS,7})^{6/5}}. \end{equation} where $M_{GMC,5}=M_c/10^5$ \mbox{M$_{\odot}$}, $n_{3}=n_{0}/10^3$ cm$^{-3}$, $c_{s,15}=c_{s}/15$ km s$^{-1}$, and $t_{MS,7}=t_{MS}/10^7$ yr. Assuming a standard IMF, this corresponds to a total star forming efficiency of about $\sim 5$ \%. For the average values of stellar ionization rates and giant molecular cloud parameters in our Galaxy, the overall star forming efficiency should be about 5\%. Obviously, larger average densities and cloud masses can result in higher star formation efficiencies. Summarizing, photoionization from OB stars can destroy the parental cloud in relatively short time scales, and defines the limiting number of newly formed stars. The fastest and most effective destruction mechanism is due to peripheral, blister, HII regions, and they can limit the star forming efficiency at galactic scales. Internal HII regions at high cloud pressures, on the other hand, result in large star forming efficiencies and they may be the main limiting mechanism in star forming bursts and at early galactic evolutionary stages (see Cox 1983). \section{Mechanical energy} As the cloud is dispersed, the average gas density decreases and the newly formed cluster becomes visible. The individual HII regions merge into a single photo-ionized structure and the whole cluster now powers an extended, low density, HII region. The stellar wind bubbles now can grow to larger sizes and some of them begin to interact. As more winds collide, the region gets pressurized by interacting winds and the general structure of the gas in the cluster is now defined by this mass and energy input (Franco {\it et al.}\ 1996). Given a total number of massive stars in the cluster, $N_{OB}$, and their average mass input rate, $<\dot{M}>$, the pressure due to interacting adiabatic winds is \begin{equation} P_i\sim \frac{N_{OB} <\dot{M}> c_i}{4 \pi r^2_{clus}}\sim 10^{-8} \frac{N_{2} <\dot{M}_6> c_{2000}}{r^2_{pc}} \ \ \ \ {\rm dyn \ cm^{-2}}, \end{equation} where $N_{2}=N_{OB}/10^2$, $<\dot{M}_6>=<\dot{M}>/10^{-6}$ M$_\odot$ yr$^{-1}$, $r_{pc}=r_{clus}/1$ pc is the stellar group radius, and $c_{2000}=c_i/2000$ km s$^{-1}$ is the sound speed in the interacting wind region. This is the central pressure driving the expansion of the resulting superbubble before the supernova explosion stage. For modest stellar groups with relatively extended sizes, like most OB associations in our Galaxy, the resulting pressure is only slightly above the ISM pressure ({\it i.e.},\ for $N_{2} \sim 0.5$ and $r_{pc}\sim 20$, the value is $P_i \sim 10^{-11}$ dyn cm$^{-2}$). For the case of rich and compact groups, as those generated in a starburst, the pressures can reach very large values. For instance, for the approximate cluster properties in starbursts described by Ho in this volume, $r_{pc}\sim 3$ and $N_{2}>10$, the resulting pressures can reach values of the order of $P_1\sim 10^{-7}$ dyn cm$^{-2}$, similar to those due to self-gravity in star forming cores. At these high pressures, the winds at the evolved red giant (or supergiant) phases cannot expand much and they reach pressure equilibrium at relatively small distances from the evolving star. Thus, the large mass ejected during the slow red giant wind phase is concentrated in a dense circumstellar shell. \begin{figure} \vspace*{43mm} \begin{minipage}{43mm} \special{psfile=fig.ps hscale=95. vscale=95. hoffset=-17 voffset=-340 angle=0. } \end{minipage} \caption{ Evolution of a wind-driven bubble in a high pressure medium. {\bf (a)} Evolution of a stellar wind from a 35\,M$_{\odot}$ star. Left scale: terminal wind velocity (km s$^{-1}$); right scale: mass loss rate (M$_{\odot}$ yr$^{-1}$); horizontal scale: time (millions of yr). {\bf (b)-(d)}: Evolution of the wind-driven bubble. The gas density (cm$^{-3}$) is plotted in a logarithmic scale against the radial distance (pc). Evolutionary times (shown in the upper-left corner of each panel) are given in million years.} \label{fig:1} \end{figure} An example of this is shown in Figure 1, where the evolution of a wind-driven bubble around a 35\,M$_{\odot}$ star is presented. Fig. 1a shows the wind velocity and mass-loss rate (dashed and solid lines, respectively: Garcia-Segura, Langer \& Mac Low 1996). We ran the simulation only over the time spanning the red supergiant and Wolf-Rayet phases, and assume that the region is already pressurized by the main sequence winds from massive stars. We used the AMRA code, as described by Plewa \& R\'o\.zyczka in this volume. During the RSG phase the wind-driven shell is located very close ($R\sim$ 0.04 pc) to the star due to a very low wind ram-pressure (Fig. 1b). Later on (Fig. 1c), the powerful WR wind pushes the shell away from the star to the maximum distance of $R\approx 0.54$ pc. Still later, when the wind has variations, the shell adjusts its position accordingly, and reaches the distance $R\sim 0.3$ pc at the end of simulation (Fig. 1d). It must be stressed that the series of successive accelerations and decelerations of the shell motion during the WR phase will certainly drive flow instabilities and cause deviations from the sphericity assumed in our model. The role of these multidimensional instabilities in the evolution of the shell is currently under study (with 2-D and 3-D models), and the results will be presented in a future communication. Regardless of the possible shell fragmentation, however, when the star explodes as a supernova, the ejecta will collide with a dense circumstellar shell. This interaction generates a bright and compact supernova remnant, with a powerful photoionizing emission ({\it i.e.},\ Terlevich {\it et al.}\ 1992; Franco {\it et al.}\ 1993; Plewa \& R\'o\.zyczka this volume), that may also be a very strong radio source, like SN 1993J (see Marcaide {\it et al.} 1995). If the shell is fragmented, the ejecta-fragment interactions will occur during a series of different time intervals, leading to a natural variability in the emission at almost any wavelength (see Cid {\it et al.} 1996). This type of interaction is also currently under investigation, and further modeling will shed more ligth on the evolution of SN remnants in high-pressure environs. This work was done during the First ``{\bf Guillermo Haro}'' Workshop, in April-May 1996, and we thank the hospitality of the staff at INAOE in Tonantzintla, Puebla. We warmly thank many useful discussions with Roberto Cid, Jos\'e Rodriguez-Gaspar, Elisabete de Gouveia, Gustavo Medina-Tanco, Michal R\'o\.zyczka, Sergei Silich, Laerte Sodre, Guillermo Tenorio-Tagle, Roberto Terlevich, and To\~ni Varela, and the enthusiasm and support given to the whole project by Alfonso Serrano, General Director of INAOE. JF and GGS acknowledge partial support to this project by DGAPA-UNAM grant IN105894, CONACyT grants 400354-5-4843E and 400354-5-0639PE, and a R\&D Cray research grant. The work of TP was partially supported by the grant KBN 2P-304-017-07 from the Polish Committee for Scientific Research. The simulations were carried out on a workstation cluster at the Max-Planck-Institut f\"ur Astrophysik. {\small
proofpile-arXiv_065-449
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} Among the most important applications of lattice QCD are the determinations of the fundamental parameters of the standard model in the quark sector. Of these, one of the most important is the overall scale of the light quark masses. It is one of the most poorly known of the parameters of the standard model from prelattice methods. The Particle Data Group estimates a range of a factor of three in allowed values \cite{RPP96}: \begin{eqnarray} 100\ {\rm MeV}&<&\overline{m}_s(1\ {\rm GeV}) <300\ {\rm MeV,} \\ 5\ {\rm MeV}&<&\overline{m}_d(1\ {\rm GeV}) <15\ {\rm MeV,\ and} \\ 2\ {\rm MeV}&<&\overline{m}_u(1\ {\rm GeV}) <8\ {\rm MeV.} \end{eqnarray} (I will use $\overline{m}_q$ to represent the running quark mass in the $\overline{MS}$ scheme.) Global fits to standard model parameters are very sensitive to such large variations. It is also one for which lattice methods provide the only systematically improvable determination. This can be contrasted with the strong coupling constant $\alpha_s$, for example, for which it is possible to imagine going to higher and higher energy scattering experiments and extracting $\alpha_s$ with perturbation theory. Lattice quark mass extractions are harder to do with solid error analysis than $\alpha_s$ extractions, but they may be more important in the long run. \section{Prelattice Quark Mass Results} \subsection{Quark mass ratios} The ratios of light quark masses can be investigated with some degree of reliability using chiral perturbation theory ($\chi$PT), which becomes asymptotically exact in the zero quark mass, zero energy limit of QCD~\cite{Leu90}. One combination of the light quark mass ratios is especially likely to be reliable, since it has been constructed to have vanishing leading order corrections in $\chi$PT: \begin{eqnarray} &&\frac{m_s^2-m_l^2}{m_d^2-m_u^2}= \frac{M_{K}^2}{M_{\pi}^2} \frac{M_K^2-M_\pi^2}{M_{K^0}^2-M_{K^+}^2+M_{\pi^+}^2-M_{\pi^0}^2} \nonumber \\ &&\times \left[1+{\cal O}(m_s^2)+{\cal O}\left(e^2\frac{m_s}{m_d-m_u}\right)\right] \end{eqnarray} The other combination of quark mass ratios may be obtained at leading order in chiral perturbation theory from the canonical prediction of $\chi$PT, that the meson mass squared is proportional to the quark mass: \begin{equation}\label{msml} \frac{m_s+m_l}{2m_l}=\frac{M_{K^0}^2}{M_{\pi^0}^2}\left[1+\ldots\right]. \end{equation} The usefulness of this relation has been called into question by the discovery of a symmetry of the chiral Lagrangian which leaves physical predictions invariant under simultaneous transformations of the interactions, and of the quark masses~\cite{kap86}: \begin{eqnarray} {\cal L}_\chi &\rightarrow& {\cal L}_\chi' \\ m_u &\rightarrow& m_u + \lambda m_s m_d, \end{eqnarray} and cyclic in u,d,s. The effects on the pseudoscalar spectrum of a nonzero $u$ quark mass can be mimicked exactly by higher order interactions in the $\chi$PT Lagrangian and altered $d$ and $s$ quark masses. In light of this, Leutwyler has assembled a variety of arguments to test the size of the corrections to Eq. (\ref{msml}) and in particular the possibility of $m_u=0$ \cite{Leu96}. He concludes that the corrections are small and obtains \begin{eqnarray} m_s/m_d &=& 18.9(8),\\ \label{sd} m_u/m_d &=& 0.553(43). \label{ud} \end{eqnarray} \subsection{Absolute value of the quark masses} Chiral perturbation theory makes no statement about the third combination of light quark masses, the overall scale. A large range of results has been obtained with a variety of methods. See, for example, a compilation of results for $m_s$ from Ref. \cite{RPP96} shown in Table \ref{tab:1}. The most reliable of these prelattice results are perhaps those using QCD sum rules, but even here, the systematic errors of the method are hard to pin down. It is here that there is the largest spread in existing results and that lattice methods probably have the most important role to play. \begin{table}[hbt] \caption[duh]{Compilation of results for $m_s$ from Ref.~\cite{RPP96}. If defined, the renormalization scheme is $\overline{m}_s(1\ {\rm GeV})$. } \begin{tabular}{l|l} \hline $m_s$ (MeV) & method \\ \hline 194 (4) & quark model \\ 118 & sum rules \\ 175 (55) & sum rules \\ $>300$ & sum rules \\ 112 (66) & $\chi$PT + estimate of $\langle \overline{q}q\rangle$ \\ 378 (220) & $\chi$PT + estimate of $\langle \overline{q}q\overline{q}q\rangle$ \\ 150 & strange baryon splittings \\ 135 & SU(6) \\ \hline \end{tabular}\label{tab:1} \end{table} \section{Lattice Quark Mass Determinations} Lattice determinations of standard model parameters require: 1) fixing the bare lattice parameters from physics, and 2) obtaining the $\overline{MS}$ parameters from these with short-distance matching calculations. Spin averaged splittings in the $\psi$ and $\Upsilon$ systems are convenient quantities to set the lattice spacing \cite{El-Khadra+al}. The mesons are small and easy to understand because the quarks are nonrelativistic. Light pseudoscalar meson masses are the most convenient quantities to fix the light quark masses. Chiral symmetry makes them very sensitive to the quark masses, the mesons are small, the correlators have good statistics and are easy to fit over long time separations. The correlators have simple behavior as the quark mass is varied toward zero, in contrast with unstable vector mesons. Other quantities must give the same results as the approximations are removed. The second piece is the determination of the parameters of the $\overline{MS}$ Lagrangian by matching perturbative, dimensionally regularized short distance amplitudes to their lattice counterparts. It is desirable to do the lattice part of such calculations nonperturbatively as much as possible, to test for the presence of nonperturbative short-distance effects and possible poor convergence of perturbation theory. Such nonperturbative short-distance calculations are harder to design for quark masses than for the strong coupling constant. Such nonperturbative short distance analysis of quark mass extractions is currently less advanced than the analogous investigations for the strong coupling constant $\alpha_s$. The perturbative expression giving the $\overline{MS}$ mass from the lattice bare mass $m_0$ may be written \begin{equation}\label{m0msbar} \overline{m}(\mu) = \tilde{m}\left[1+g^2 \left( \gamma_0 \left( \ln \tilde{C}_m-\ln\left( a \mu\right) \right) \right) \right]. \label{eq:star} \end{equation} The mean-field-improved bare mass $\tilde{m}$ is given by $\tilde{m} = m_0/(1-\frac{1}{12}g^2)$ in perturbation theory, and $\tilde{m}=m_0/\sqrt[4]{\langle U_P\rangle_{MC}}$ nonperturbatively, if the expectation value of the plaquette, $\langle U_P\rangle$, is used to define mean field improvement\cite{LM93}. $\gamma_0=1/(2\pi^2)$ is the leading quark mass anomalous dimension. This coefficients in this expression are well behaved for the Wilson action and the ${\cal O}(a)$ improved action of Sheikholeslami and Wohlert (SW) \cite{She85} (for which $\tilde{C}_m = 1.67$ \cite{Gro84} and 4.72 \cite{Gab91} respectively). For staggered fermions, $\tilde{C}_m$ is 132.9 \cite{Gro84}, leading to renormalization factors of 50--100\% which are not explained by large tadpole graphs. The status of quark mass extractions as of a few years ago was reviewed in Refs.~\cite{Uka93,Gup94}. Some data tabulated in Ref.~\cite{Uka93} are shown in Fig.~\ref{old}. Unimproved perturbation theory was used, and the lattice spacing was determined with the $\rho$ mass. The quenched results for staggered fermions (dark squares) are relatively independent of the lattice spacing, but Eq.~\ref{m0msbar} is untrustworthy because of the very large quantum correction. The quenched Wilson results (diamonds) have much better controlled perturbation theory, but are much more lattice spacing dependent. If the results are extrapolated in $a$, they approach the staggered results more closely. However, the magnitude and origin of remaining lattice spacing errors is then unknown. Improved actions must be used to investigate this. \begin{figure} \epsfxsize=0.45 \textwidth \epsfbox{oldcombined.eps} \caption[old]{ Old lattice quark mass results were reviewed in Ref.~\cite{Uka93}. Quenched staggered results (dark squares) show good cutoff independence, but suffer from huge perturbative corrections. Quenched Wilson results (circles) had well-converged perturbation theory, but large cutoff dependence. The white squares show some unquenched staggered results. } \label{old} \end{figure} \section{Recent Quenched Lattice Results} \subsection{Unimproved Wilson fermion results.} New unimproved results have been presented in the JLQCD contribution to this volume by T. Yoshi\'e \cite{yos96}. They include a data point at $\beta=6.3$, corresponding to a smaller lattice spacing ($a=(3.29\ {\rm GeV})^{-1}$) than previous results. Fig. \ref{jlqcd} shows their Wilson fermion results superimposed on a subset of the world data. $M_\rho$ has been used to set the lattice spacing. Improved perturbative theory has been used in the renormalizations, which makes the Wilson results slightly higher than in the older analyses, and the staggered results much higher, about 50\%. The new JLQCD Wilson results are statistically consistent with previous results, and so can be combined with them for a consistent extrapolation of the leading error. Taken by themselves, however, they appear to have a somewhat smaller $a$ dependence then other results. Wilson data were also presented by Fermilab (upper points in Fig.~\ref{fermilab}) \cite{Ono96,gou96}. These points also lie slightly below and have a smaller slope than most of the world data in Fig.~\ref{jlqcd}. This is due to the fact that the Fermilab data use the spin averaged 1P--1S splitting in the $\psi$ system to determine the lattice spacing, which is expected to have a small ${\cal O} (a)$ error. Determining $a$ with the $\rho$ mass is equivalent to calculating $m_l/m_\rho$. Since the $ \overline{\psi} \sigma_{\mu\nu} F_{\mu\nu} \psi$ ${\cal O} (a)$ correction operator is expected to push spin partners like the $\pi$ and the $\rho$ apart, ${\cal O} (a)$ errors in both masses contribute to the $a$ dependence. When the 1P--1S splitting is substituted for the $\rho$, the slope ought to be reduced, as observed. If $\rho$ mass lattice spacings are substituted into the Fermilab data, the results line up with the upper points in Fig.~\ref{jlqcd}. \begin{figure} \epsfxsize=.45 \textwidth \epsfbox{fig-3a.ps} \caption[jlqcd]{ Recent unimproved Wilson fermion results for $\overline{m}_l(2\ {\rm GeV})$ from JLQCD compared with previous world data (upper points) \cite{yos96}. Lattice spacing from $M_\rho$. JLQCD staggered fermion results are the lower $*$'s. } \label{jlqcd} \end{figure} Gupta and Bhattacharya have performed a linear extrapolation on the world Wilson data and obtain $\overline{m}_l(2\ {\rm GeV})= 3.1(3)$ MeV \cite{Gup96}. They use slightly different analysis methods from those of Ref.~\cite{yos96}, and obtain a result below that which would be obtained from Fig.~\ref{jlqcd}, which is about 3.8 MeV. However, the authors say that their error estimate is a ``guess chosen to reflect the various uncertainties discussed \dots and is not based on systematic analysis''. JLQCD also presented new results for staggered fermions which left the existing situation unchanged. The results are almost lattice independent for $a<(1.5\ {\rm GeV})^{-1}$, but it is difficult to estimate perturbative uncertainties since the perturbative renormalization factor is so large. (At $\beta=6.4$, they use $Z=1.69$, a 69\% correction of which most is unexplained by mean field improvement.) \subsection{Improved quenched Wilson fermion results} To investigate the size of the remaining cut--off errors in the Wilson action (${\cal O} (a^2)$, ${\cal O} (a \alpha_s)$, ${\cal O} (\alpha_s^2)$, etc.), it is necessary to remove the leading error with an improved action. In the SW action, the leading ${\cal O} (a)$ error is removed with the addition of the operator $ \overline{\psi} \sigma_{\mu\nu} F_{\mu\nu} \psi$. There are several gauge links in the operator, so tadpole improvement predicts a rather large correction, of order 50\%. New results with the tadpole improved SW action were presented by Fermilab \cite{Ono96}. The lattice spacing dependence observed in the Wilson points is substantially reduced, but not completely eliminated. If this arises from a residual ${\cal O} (a)$ error, the results extrapolate down to rather near the staggered results. (A recent nonperturbative determination of the coefficient of the SW improvement operator has indicated the possibility of even larger corrections than the sizable ones indicated by perturbation theory and mean field improvement~\cite{Jan96}.) If it arises from higher order effects in $a$, the answer is close to the existing small lattice spacing points, and the discrepancy with staggered results must be attributed to poorly behaved staggered perturbation theory. \begin{figure} \epsfxsize=0.45 \textwidth \epsfbox{mq.eps} \caption[fermilab]{ Fermilab light quark masses for Wilson action (circles) and the tadpole improved SW action (triangles)~\cite{Ono96}. Lattice spacing from the 1P--1S splitting in the $\psi$ system. } \label{fermilab} \end{figure} Since a downward trend in $a$ is still present in the tadpole improved data, we take the improved result at the finest lattice spacing as an upper bound in the quenched approximation and a linear extrapolation as a lower bound. Adding perturbative uncertainties linearly and other uncertainties in quadrature gives the quenched result: \begin{eqnarray}\label{eq:qu} \overline{m}_l(2\ {\rm GeV}) & = & 3.6 (6)\ {\rm MeV} ,\\ \overline{m}_s(2\ {\rm GeV}) & = & 95 (16)\ {\rm MeV}. \end{eqnarray} Another determination of the strange quark mass with an ${\cal O}(a)$ improved action has been reported, in Ref. \cite{All94}. This determination used a tree-level, rather than a mean-field improved coefficient for the improvement operator. They obtained $\overline{m}_s(2\ {\rm GeV})=128 \pm 18$ MeV. They did not attempt to correct for the effects of the remaining lattice spacing dependence or the effects of the quenched approximation. Much of the discrepancy with the Fermilab results arises from fact that we have used much larger SW improvement coefficients, and make an allowance for the fact that we continue to find significant cut-off dependence even so. \section{Corrections to Quark Mass Ratios} \subsection{Nonlinearities in $m_q$ vs. $M_\pi^2$} \begin{figure} \epsfxsize=0.45 \textwidth \epsfbox{aoki.ps} \caption[nonlin]{ $(aM_\pi)^2$ vs. $am_q$ for staggered fermions at $\beta=6.4$ on a $40^3\times 96$ lattice \cite{jlq96}. Deviations from linearity are a few \% or less. } \label{nonlin} \end{figure} There is very little lattice evidence for large deviations from linearity in the quark masses in Eq. (\ref{msml}) in either quenched or unquenched calculations. In the most accurate data, the question often seems to be why the predictions of chiral symmetry appear linear at the few~\% level up to such large values of the quark masses. (See Fig.~\ref{nonlin}. The largest pion mass is around 850 MeV, and nonlinearity in $(aM_\pi)^2$ vs. $am_q$ is only a few~\%.) However, searching for such deviations is tricky, especially in the quenched approximation where quenched chiral logarithms may add spurious nonlinearities at small $m_q$ \cite{Sha96}. Existing unquenched calculations have not yet examined carefully the case of broken flavor SU(3), $m_s>m_l$, so not all of the higher order operators of the chiral Lagrangian have yet been tested. \subsection{Electromagnetic effects} It is in principle simple to determine the contributions of electromagnetism to hadron masses using numerical simulations with link matrices which are products of SU(3) and U(1) matrices. The U(1) phases of electromagnetism are cheap to generate because to leading order in $\alpha_{em}$ (all that is required for practical purposes), they can be obtained from Fourier transforms of Gaussian fluctuations in momentum space. To see the effects of electromagnetism clearly above fluctuations in the SU(3) field, one wants to use values of $\alpha_{em}$ which are larger than the physical value. It then remains only to show that electromagnetic effects are still linear in this region. That this indeed holds has been shown recently by Duncan, Eichten, and Thacker \cite{dun96}. They have performed a prototype calculation of the $\pi^+ - \pi^0$ splitting at $\beta=5.7$ and with it have obtained $m_u/m_d = 0.51$, in agreement (so far) with Eq.~\ref{ud}. \subsection{Can $m_u = 0$?} The most interesting application of these calculations is the settling of the question of whether $m_u \equiv 0$ in the real world. This possibility is fervently desired in spite of all evidence to the contrary because of its neat solution to the strong CP problem. Ref.~\cite{kap86} showed that allowing corrections as large as 30\% to $\chi$PT equations such as \ref{msml} made $m_u = 0$ compatible with meson mass data. (See Fig.~\ref{nonlin}, however.) The possibility of large instanton-induced flavor mixing effects has been proposed as a mechanism to generate such corrections in QCD~\cite{Geo81}. Not all the required lattice calculations have been done, but so far, the lattice evidence is against $m_u=0$. \section{The Quenched Approximation}\label{quapp} As the lattice spacing is reduced, while keeping hadronic physics fixed, the lattice bare couplings evolve according to their anomalous dimensions. In the quenched approximation, these anomalous dimensions are slightly wrong, due to the absence of light quark loops. The strong coupling constant evolves according to the zero-quark $\beta$ function coefficient $\beta_0^{(0)}=11$ rather than the correct three-quark coefficient $\beta_0^{(3)}=11-2/3 n_f=9$. Asymptotically, $\alpha_s(\pi/a)$ in the quenched approximation is expected to be too small by a factor of about 9/11. Since the short-distance quark mass evolution is given by $d\ln m(q)/d \ln(q) = - \gamma_0 \alpha_s/(4\pi)$, where $\gamma_0 = 8$, this implies that the quark mass evolves too slowly in the quenched approximation, and at small lattice spacings is larger than in real life. At high energies, running mass evolution is given by \begin{equation} \frac{m(q_1)}{m(q_2)} \approx \left( \frac{\alpha_s(q_1)} {\alpha_s(q_2)}\right)^{\frac{\gamma_0}{2\beta_0}}. \end{equation} To leading log accuracy, therefore, the effect of the absence of quark loops due to {\em perturbative} effects on the evolution of the running mass from the strong coupling region (where $\alpha_s\approx 1$) to the high energy region can be approximated by \cite{Mac94} \begin{eqnarray} \frac{m(\pi/a)|_{\rm qu.\ \ \ }}{m(\pi/a)|_{\rm unqu.}} &\approx& \alpha(\pi/a)^{\frac{\gamma_0}{2}(1/\beta_0^{(0)}-1/\beta_0^{(3)})}\\ &\approx& 1.1 {\ \rm to \ } 1.2, \label{eq:1.15} \end{eqnarray} for $\alpha(\pi/a)\approx 1/4$ to 1/8. There are also quenching effects arising from the nonperturbative region. Unlike the case of quarkonium systems, for pions there is no argument that these should be smaller than the perturbative effects. The above expression can't be taken as a correction factor, only as an indication of the direction and order of magnitude of effects to be expected. Nonperturbative calculations are required to investigate quenching effects quantitatively. Unquenched results for Wilson fermions appear complicated and hard to interpret. They differ by as much as a factor of two from quenched results, and do not seem to reproduce the lattice spacing dependence of the quenched results as would have been expected (see Ref. \cite{Uka93}). Since most other unquenched Wilson results, such as those in thermodynamics, are also hard to understand compared to staggered results, I will not attempt to fit unquenched Wilson fermion results into my general picture. Some unquenched staggered results summarized in Ref. \cite{Uka93} are shown in Fig. \ref{old} (white squares) along with the quenched results. The unquenched results indeed lie below the quenched results by roughly the expected amount, and I will use them to estimate the effects of quenching. The large corrections in staggered fermion mass renormalization cancel out in the ratio of the quenched and unquenched determinations, making this an useful quantity to examine. To minimize effects due to differences in analysis methods, we estimate the ratio from the results of a single group, at similar volumes and lattice spacings (about 0.4 GeV$^{-1}$) \cite{Ish92,Fuk92}, and obtain \begin{eqnarray} \frac{\overline{m}_l(1.0\ {\rm GeV})_{\rm n_f=0}}{\overline{m}_l(1.0\ {\rm GeV})_{\rm n_f=2}} &\approx& \frac{2.61(9)}{2.16(10)}\\ &=& 1.21(7) \label{eq:unq} \end{eqnarray} Since there are, in fact, three flavors of light quarks in the world and not two, I will this ratio as a lower bound on the actual ratio and use the square (corresponding to four light quarks) as an upper bound. \section{Synthesis} Lattice determinations of light quark masses are more difficult than the analogous determinations of $\alpha_s$. Pion masses have worse $a$ dependence than the quarkonium splittings. Finding nonperturbative ways of eliminating the perturbative relations with the bare lattice quark mass requires more work. Nevertheless, with a few exceptions, most lattice quark mass extractions are consistent with a reasonably simple picture. I ignore results with very large $a$ or very small volume. Unquenched Wilson results, which are also hard to interpret in most other quantities, do not seem to make sense compared with quenched Wilson results. Of the remaining results, the larger magnitude and lattice spacing dependence of Wilson results compared with staggered results is greatly reduced with ${\cal O} (a)$ improved actions. The magnitudes are reduced more with tadpole improved SW correction terms than with tree--level correction terms. If the remaining discrepancy arises mainly from residual ${\cal O} (a)$ effects in the improved action, the true quenched answer lies close to the staggered result, $\overline{m}_l$ a little over 3 MeV. If it arises mainly from higher order corrections in the staggered fermion lattice--$\overline{MS}$ mass conversion (where the leading correction is 50--100\%), the true answer lies closer to the improved result, $\overline{m}_l \sim 4$ MeV. Unquenched staggered results lie somewhat below quenched staggered results, but by an amount which is reasonable. Taking the ratio from Sec.~\ref{quapp} and using it to make a correction on our quenched result we obtain \begin{itemize} \item $\overline{m}_s(2\ {\rm GeV})$ in the range 54 -- 92 MeV, \item $\overline{m}_l(2\ {\rm GeV})$ in the range 2.1 -- 3.5 MeV, \end{itemize} for the $\overline{MS}$ masses renormalized at 2 GeV. The uncertainties most in need of further study are those associated with lattice spacing dependence and the quenched approximation. \section*{Acknowledgments} I thank Brian Gough, Aida El-Khadra, George Hockney, Andreas Kronfeld, Bart Mertens, Tetsuya Onogi, and Jim Simone for collaboration on the work of Ref. \cite{Ono96,gou96}. I thank the Center for Computational Physics in Tsukuba for hospitality while this paper was written, and I thank the members of the JLQCD collaboration for useful discussions. Fermilab is operated by Universities Research Association, Inc. under contract with the U.S. Department of Energy.
proofpile-arXiv_065-450
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{...} is the same as in LaTeX \def\section#1{\ifblank\sectionnumstyle\then \else\newsubsectionnum=\z@\@resetnum\sectionnum=\next \fi \displayhead{\ifblank\sectionnumstyle\then\else\sectionnumformat\ \fi#1} } \def\subsection#1{\advance\subsectionnum by 1 {\boldfont \leftline{\ifblank\sectionnumstyle\then\else\sectionnumformat\fi\number\subsectionnum\ #1}} \vskip-10pt \line{\hrulefill}} \def \section{Introduction} \medskip\noindent Dynamical systems that are naturally close to or on their critical point have been proposed as an explanation of the appearance of power laws in nature \cite{1,2}. This `self-organized criticality' still consists to a large extent in the study of toy models and has many open or controversial questions. The lattice models can be grouped at least into three classes. A first class contains models with a local conservation law like the famous sandpile-model \cite{1,2}. Recent experiments on piles of rice \cite{3} partially confirm the general theoretical predictions by exhibiting power laws, but also show that the existence of power laws in real systems depends on microscopic details such as the aspect ratio of grains of rice. Models of evolution \cite{4,5} are one example in a second class where the dynamics is specified in terms of a globally selected extremal site. This extremal dynamics can be used to obtain some general statements for the complete class of models \cite{6}. A model for self-organized criticality that is known as the `forest-fire model' is a member of a third class of models that have parameters which can be tuned close to the critical point in a natural manner, but unlike in the two previous classes cannot be entirely discarded (In dealing with this model one should be aware that it is highly idealized and not expected to describe real forest fires). The precise version that we study in this paper has first been introduced in a short note \cite{7} and arises as a certain limit of the more general model proposed later independently in \cite{8}. \medskip\noindent The two-dimensional forest-fire model has been already discussed very controversially, mainly on the basis of Monte-Carlo simulations where usually the accuracy of the predictions was the issue. An originally proposed version \cite{9} did not show the desired critical behaviour \cite{10,11}, and it was necessary to introduce lightnings \cite{8}. Subsequently, Monte-Carlo simulations have several times lead to values for the critical exponents which had to be corrected later on \cite{8,12,13,14}. Here we add to this discussion by re-examining some of the quantities investigated only in one of the earlier works \cite{12}. We were motivated by a study of the one-dimensional case \cite{15} where we had discovered the existence of two different length scales. The analogous question in two dimensions has been addressed in \cite{12}, but there it seemed that the two scales are proportional to each other. Here, we present more accurate simulations that demonstrate these two length scales to be different also in two dimensions. As by-products we also check or improve estimates for other exponents (see in particular \cite{14}). A final subject is the approach to equilibrium which seems not to have been addressed in a similar way before. In a second part, we try to discard the spatial structure and introduce a global model similar to the one of \cite{15} for one dimension. On the one hand, some of the qualitative features of the stationary state at the critical point (e.g.\ the power law of the cluster-size distribution) can nicely be described by such a global model. On the other hand, there are discrepancies in quantitative details and the range of such a simplified model is very limited in comparison to the full model. We believe that this is an important point, e.g.\ because it has been suggested in \cite{16} that power laws in nature might arise from a global (`coherent') driving that does not see any spatial structure. Our findings here and in the one-dimensional case \cite{15} demonstrate that the full model is not only different from but also richer than the simplified model. This is analogous to the result of \cite{17} that versions of certain models of evolution with and without spatial structure lead to different results (at least if examined closely). \medskip\noindent We now define the model before we proceed with a presentation of our simulation results in the next Section. The forest-fire model is defined on a cubic lattice in $d$ dimensions. Any site can have two states: It can either be empty or it can be occupied by a tree. The dynamics of the model is specified by the following update rules (following \cite{7,12,13,14} \footnote{${}^{1})$}{ To be precise, the simulations in \cite{13,14} have been performed according to slightly different rules, because they aimed at investigating only quantities associated with clusters. }): In each Monte-Carlo step first choose an arbitrary site of the lattice. \item{a)} If it is empty, grow a tree there with probability $p$. \item{b)} If it is occupied by a tree, delete the entire geometric cluster of trees connected to it with probability $f$. This corresponds to a lightning stroke with subsequent spreading of the fire. \par\noindent A rescaling of the probabilities $p$ and $f$ just amounts to a rescaling of the time scale, and in particular leaves the stationary state invariant. We exploit this to set $p=1$. There is a critical point at $f/p = 0$, but the parameter $f/p$ is relevant and it is not legitimate to consider the forest-fire model precisely at this critical point (compare \cite{18}). \bigskip\noindent \section{Simulation results in two dimensions} \medskip\noindent In the following we consider the two-dimensional version of the forest-fire model on a quadratic lattice with periodic boundary conditions. The linear size of the lattice will be denoted by $L$. So the volume $V$ is given by $V = L^2$. We use a `global' time scale, a unit of which is defined by the number of Monte-Carlo steps needed in order to visit each site on average once, i.e.\ a unit of global time consists of $L^2$ Monte-Carlo steps. \medskip\noindent In order to do the simulation efficiently also for large systems and small $f/p$ one has to be careful and use e.g.\ bitmapping technologies. For details on the implementation compare the WWW page \cite{19}. Using this program, the simulations of this paper took about four months of CPU time on 150MHz DEC alpha workstations. \medskip\noindent We have investigated mainly systems of linear size $L=16384$ and parameter values $10^{-2} \ge f/p \ge 10^{-4}$. For a simulation, one random initial condition with density $\rho = 1 / 2$ was chosen. In order to equilibrate the system, it was left to evolving freely for at least 15 global time units. This equilibration time was increased to 25 global time units for $f/p \le 3 \cdot 10^{-4}$ and to 35 global time units for $f/p = 1 \cdot 10^{-4}$ (these times were adjusted according to the observed time evolution of $\rho$, see also Section 2.3 below). After this, the system was iterated further for another 60 to 90 global time units (120 units for $f/p = 3 \cdot 10^{-4}$). During this period, measurements were made as global averages at intervals of usually one global time unit (for the density 200 times more frequently). This amounts to at least $60 L^2$ measurements for each quantity of interest. Note that we use only a single run. \bigskip\noindent \subsection{Correlation functions} \medskip\noindent Let us first discuss the simulation results for the usual tree correlation function $\langle T({\vec{x}}) T({\vec{x}}+{\vec{y}})\rangle$ ($T({\vec{x}}) = 1$ in a configuration with a tree at site ${\vec{x}}$, $T({\vec{x}}) = 0$ if the site ${\vec{x}}$ is empty). We have only investigated displacements ${\vec{y}}$ along the vertical axis (which can be treated particularly efficiently using bitmaps). The treatment of the data resulting from the simulation is straightforward because it can nicely be fitted by $$C(y) := \langle T({\vec{x}}) T({\vec{x}}+y \vec{e}_2)\rangle - \langle T({\vec{x}}) \rangle^2 = a e^{-{y / \xi}} \label{defTp}$$ in a suitable interval $y_{\rm min} \le y \le y_{\rm max}$. The parameters $a$ and $\xi$ were estimated by taking the logarithm of the r.h.s.\ of \ref{defTp} and then performing a linear regression for $y_{\rm min} \le y \le y_{\rm max}$. These bounds are chosen such that the approximation \ref{defTp} by a single exponential function is good. The lower cutoff can be chosen small ($y_{\rm min} \approx 20$) independent of $f/p$. An upper cutoff $y_{\rm max}$ has to be imposed at distances of 4 to 7 times the correlation length $\xi$ because then statistical errors become large. \medskip\noindent \centerline{\psfig{figure=length_n.ps}} \smallskip\noindent {\par\noindent\leftskip=4.5 true pc \rightskip=4 true pc {\bf Fig.\ 1:} The two correlation lengths $\xi$ and $\xi_c$. For $\xi_c$ all estimates are on lattices with $L=16384$ and the different symbols correspond to different ways of treating the data. \par\noindent} \medskip\noindent Fig.\ 1 shows the results for $\xi$ obtained in this manner on lattices with $L=8192$ (diamonds), $L=16384$ (crosses) and $L=17408$ (triangles). One observes that they are in good agreement with the form $$\xi \sim \left({f \over p}\right)^{-\nu_T} \, . \label{DEFnuT}$$ Performing linear regression fits on a doubly logarithmic scale one finds \footnote{${}^{2})$}{ Error estimates are always the $1 \sigma$ confidence interval of a fit unless discussed explicitly.} $$\nu_T = 0.541 \pm 0.004\, . \label{RESnuT}$$ This is close to the result $\nu_T = 0.56$ found in \cite{12}. Comparing the values for $\xi$ obtained from simulations on different lattice sizes one can see that they may have a residual statistical error of around 2\% which is about the same as the scattering of the individual data points around the line \ref{DEFnuT}. Since also further data for larger $f/p$ is consistent with the form \ref{DEFnuT} and the value \ref{RESnuT}, this result for $\nu_T$ and its error bound can be considered reliable. \medskip\noindent The estimates for the normalization constant $a$ in \ref{defTp} are compatible with a value $a = 0.030 \pm 0.001$ independent of $f/p$. So, in the limit $f/p \to 0$ the two-point function seems to tend to $\langle T({\vec{x}}) \rangle^2 + a$ for $y \age 100$. In terms of the alternative ansatz $C(y) = a y^{-\eta_{{\rm occ}}} e^{-{y / \xi}}$ used in \cite{12} this corresponds to $\eta_{{\rm occ}} = 0$, while the result found there was $\eta_{{\rm occ}} = 0.120 \pm 0.015$. We have also looked into the possibility of a power-law correction factor using our data for $L=16384$ and $f/p \le 3 \cdot 10^{-4}$. With the estimates for $\xi$ shown in Fig.\ 1 one finds that $e^{{y / \xi}} C(y) \sim y^{-0.11}$ for $y \le 20$, i.e.\ for small $y$ there is indeed a power-law correction factor with an exponent that is consistent with \cite{12}. On the other hand, for $50 \le y \le 4 \xi$, the function $e^{{y / \xi}} C(y)$ is flat -- its smallest values are around $0.027$ and its maximal values around $0.031$. This clearly contradicts a power-law correction factor with $\eta_{{\rm occ}} \approx 0.11$ in the large-distance asymptotics. Thus, for the large-distance behaviour $\eta_{{\rm occ}} = 0$ seems to be correct, while the power-law correction factor observed in \cite{12} applies to small distances. \medskip\noindent We now look at a second quantity, namely the `connected correlation function' $\langle T({\vec{x}}) T({\vec{x}}+{\vec{y}})\rangle_c$ describing the probability to find two trees at positions ${\vec{x}}$ and ${\vec{x}}+{\vec{y}}$ {\it inside the same cluster}. This quantity is usually referred to as {\it the} two-point function in the context of self-organized criticality and often is also the only correlation function that is investigated. We demonstrated in one spatial dimension \cite{15} that the length scales associated to this correlation function and $C(y)$ have different critical exponents while \cite{12} suggested that in two dimensions length scales associated to different quantities are equivalent. This latter suggestion was based on simulations with lattice sizes up to $512 \times 512$ and $f/p \age 5 \cdot 10^{-4}$. One of the main aims of the simulations presented here is to check if different length scales can be exhibited also in two dimensions after sufficiently improving the accuracy. As before, we restrict to displacements ${\vec{y}}$ along the vertical axis, i.e.\ we consider $$K(y) := \langle T({\vec{x}}) T({\vec{x}}+y \vec{e}_2)\rangle_c \, . \label{DEFconTPF}$$ This quantity can be determined in the same run as the two-point function $\langle T({\vec{x}}) T({\vec{x}}+{\vec{y}})\rangle$, but it involves the additional effort of determining all clusters present in the system. \medskip\noindent Following \cite{14} we associate a correlation length $\xi_c$ to the second moment of $K(y)$ via $$\xi_c^2 := {\sum_{y=1}^{\infty} y^2 K(y) \over \sum_{y=1}^{\infty} K(y)} \, . \label{DEFxiC}$$ The squares in Fig.\ 1 show values of $\xi_c$ extracted from simulations with $L=16384$ using this definition. One sees that these values are consistent with $$\xi_c \sim \left({f \over p}\right)^{-\nu} \, . \label{DEFnuC}$$ and a (preliminary) value of $$\nu = 0.5738 \pm 0.0013 \, . \label{RESnuCpre}$$ This agrees roughly with the value $\nu = 0.60$ found in \cite{12}, and also with the more accurate result in \cite{14} which, including error bounds, is given by $\nu = 0.580 \pm 0.003$ \cite{20}. It should be noted that the lower bound $y=1$ in \ref{DEFxiC} is crucial. Starting the summation instead e.g.\ at $y=10$, one finds $\xi_c \approx 24$ at $f/p = 10^{-2}$ and $\xi_c \approx 219$ for $f/p = 10^{-4}$ instead of $\xi_c \approx 14$ and $\xi_c \approx 203$, respectively. This in turn would make the value for $\nu$ smaller. We will argue soon that the start of summation $y=1$ is indeed the correct choice, but the impact of a modification here could give rise to a slightly larger error than the $1\sigma$ interval for the fit given in \ref{RESnuCpre}. \medskip\noindent We now proceed with a more detailed discussion of the form of $K(y)$ which will also justify the definition \ref{DEFxiC}. Eq.\ \ref{DEFxiC} is based on the following expected form of $K(y)$: $$K(y) = a_c \; y^{-\eta} \; e^{-y/\xi_c} \, . \label{DEFeta}$$ In particular, if $e^{y/\xi_c} \; K(y)$ agrees well with a power law, the determination of $\xi_c$ via \ref{DEFxiC} is justified. Using the values of $\xi_c$ given by the boxes in Fig.\ 1 one finds that $e^{y/\xi_c} \; K(y)$ does indeed agree well with $a_c \; y^{-\eta}$ for $y$ between 1 and several times $\xi_c$ \footnote{${}^{3})$}{ The values of $a_c$ are all compatible with the $f/p$-independent value $a_c = 0.20 \pm 0.01$. Therefore, at the critical point the power law $K(y) = a_c y^{-\eta}$ is expected to be valid for all $y$.}. The crosses in Fig.\ 2 show these estimates for $\eta$. They obviously depend on $f/p$ contrary to what one would naively expect. This points to systematic errors in the determination of $\eta$ which one may expect to become less important for smaller $f/p$ where the power law is better visible. This suggests to make a `scaling ansatz' $\eta(f/p) = \eta + \bar{a} \; (f/p)^{\bar{b}}$ to determine the value of $\eta$ in the limit $f/p \to 0$. A least-squares fit gives the dotted line in Fig.\ 2. One finds a critical $\eta = 0.374 \pm 0.011$. \medskip\noindent \centerline{\psfig{figure=eta.ps}} \smallskip\noindent {\par\noindent\leftskip=4.5 true pc \rightskip=4 true pc {\bf Fig.\ 2:} Values for the exponent $\eta$ obtained in two different ways (crosses and diamonds) and scaling fits (lines). \par\noindent} \medskip\noindent Alternatively, one can fit all three parameters in \ref{DEFeta} directly from the data. The values for $\eta$ obtained with this approach are shown in Fig.\ 2 by diamonds. A scaling analysis (the line with long dashes in Fig.\ 2) yields a critical $\eta = 0.342 \pm 0.008$. Comparing this with the value found earlier, one finds a small disagreement. So, the error bounds of these two estimates for the critical $\eta$ may be a little too optimistic because they do not include systematic errors. In order to be careful we give a final result that includes both estimates for $\eta$ and the two error ranges: \goodbreak $$\eta = 0.36 \pm 0.03 \, . \label{RESeta}$$ This is compatible with the result $\eta = 0.411 \pm 0.02$ of \cite{12} (although the error estimate of \cite{12} seems a little too optimistic, possibly because it does not take systematic errors into account). One also obtains different estimates for $\xi_c$ from the same fits that gave the alternative values for $\eta$. These values are shown in Fig.\ 1 by `$\times$'. One sees that they are again compatible with the form \ref{DEFnuC}, and obtains an alternative estimate for the associated critical exponent: $\nu = 0.5777 \pm 0.0013$. There is a minor difference between this value and the earlier estimate \ref{RESnuCpre} indicating that we have indeed neglected systematic errors. Therefore we give a final result $$\nu = 0.576 \pm 0.003 \label{RESnuC}$$ which includes the two direct estimates and their error bound. This final result is in excellent agreement with \cite{14}. \medskip\noindent Comparing \ref{RESnuC} with \ref{RESnuT}, we see that $\nu \ne \nu_T$, so that $\xi_c$ and $\xi$ are basically different lengths. A difference $\nu - \nu_T \approx 0.04$ was already observed in \cite{12}, but attributed to numerical errors because it was unexpected. \medskip\noindent It should be mentioned that our results \ref{RESeta} and \ref{RESnuC} for $\eta$ and $\nu$ do not satisfy the scaling relation $2-\eta = 1/\nu$ \cite{12}. We can only speculate about the reason for the disagreement. For example, in the derivation of this scaling relation one assumes that $\eta$ and $\nu$ do not depend on the direction of the displacement vector ${\vec{y}}$, and we have not checked whether this is indeed true. Another possibility is that we have still overlooked systematic errors. Inserting $\nu$ (which is the more reliable value) according to \ref{RESnuC} into the scaling relation $2-\eta = 1/\nu$ yields $\eta \approx 0.28$ which is possible if our error estimate in \ref{RESeta} is by a factor of about 3 too small. \bigskip\noindent \subsection{Cluster-size distribution} \medskip\noindent The distribution $n(s)$ of clusters with size $s$ arises as a by-product of the determination of $K(y)$ during the simulations. We extract an exponent $\tau$ from it following the lines described in detail in \cite{12,13}. One introduces the quantity $P(s) = \sum_{s' > s} s' n(s')$. Assuming that it behaves as $P(s) = \alpha s^{2-\tau} e^{-s/s_{{\rm max}}}$, one can use three-parameter fits to obtain estimates for $\tau$ and $s_{{\rm max}}$. Extrapolation of the values obtained in this manner from the simulations with $L=16384$ yields $\tau = 2.1595 \pm 0.0045$ for $f/p \to 0$. The values of $\tau$ for $f/p > 0$ approach this limiting value from above. An alternative way to extract a value of $\tau$ is to look directly at $n(s)$ and assume that $n(s) \sim s^{- \tau}$ for intermediate $s$. Applying this second method to the same data and to some results for $L=8192$ we obtain the estimate $\tau = 2.159 \pm 0.006$ at the critical point. This limit is now approached from below and is in excellent agreement with the one obtained before. To be on the safe side we retain the value with the larger error bound as the final result: $$\tau = 2.159 \pm 0.006 \, . \label{REStau}$$ This value agrees within error bounds with most previous results \cite{12,21,13,14}, but our error bound is considerably smaller than in most of them. \medskip\noindent When one extracts estimates for $\tau$ from $P(s)$ one also obtains estimates for $s_{{\rm max}}$. These values are in good agreement with the form $s_{{\rm max}} \sim (f/p)^{-\lambda}$ with $\lambda = 1.171 \pm 0.004$. To check the validity of our error estimate we can use the scaling relation $\lambda = 1/(3-\tau)$ \cite{14}. After inserting \ref{REStau} one finds the prediction $\lambda = 1.189 \pm 0.008$. This prediction agrees roughly with the direct estimate. However, there is a small discrepancy indicating that the estimate $$\lambda = 1.17 \pm 0.02 \label{RESlam}$$ is more realistic. Within error bounds we find good agreement with the values of \cite{14} and \cite{12}. From the results \ref{RESnuC} and \ref{RESlam} we obtain the fractal dimension $\mu = 2.03 \pm 0.04$ using the scaling relation $\mu = \lambda/\nu$ \cite{14}. This does not quite agree with \cite{14} where $\mu = 1.96 \pm 0.01$ was found, but would favour instead the expectation $\mu = 2$ of \cite{13}. Unfortunately, with our simulations we had not aimed at determining $\mu$ and we are therefore not able to clarify this interesting point. \bigskip\noindent \subsection{Density and time evolution} \medskip\noindent Here we study the critical behaviour of the stationary density as well as the temporal behaviour of the density. The latter yields the lowest gap in the spectrum of the time-evolution operator which is given by a straightforward generalization of eq.\ (3.6) in \cite{15}. Some high relaxational modes of this time-evolution operator can be written down explicitly as we demonstrate in Appendix A. In one dimension, the low-lying spectrum of this operator could be studied numerically \cite{15}, but this is not feasible in higher dimensions. Fortunately, simulations of $\rho(t)$ exhibit much clearer features in two dimensions than was the case in one dimension. This is illustrated by Fig.\ 3 which shows the initial time evolution of $\rho(t)$ after the system was started at $\rho(0) = 1/2$. \medskip\noindent \centerline{\psfig{figure=rho_t.ps}} \smallskip\noindent {\par\noindent\leftskip=4.5 true pc \rightskip=4 true pc {\bf Fig.\ 3:} The density $\rho(t)$ as a function of global time $t$. Shown are results from simulations with $f/p = 10^{-2}$ (bottom, full line), $f/p = 10^{-3}$ (line with long dashes, middle) and $f/p = 10^{-4}$ (top line with short dashes). \par\noindent} \medskip\noindent First we examine the critical behaviour of the stationary density of trees $\rho(\infty)$. The values in Table 1 are obtained by averaging $\rho(t)$ over times $t$ after the equilibration time, i.e.\ at or beyond the right border of Fig.\ 3. \medskip\noindent \centerline{\vbox{ \hbox{ \vrule \hskip 1pt \vbox{ \offinterlineskip \defheight2pt&\omit&&\omit&&\omit&\cr{height2pt&\omit&&\omit&&\omit&&\omit&\cr} \def\tablerule{ height2pt&\omit&&\omit&&\omit&\cr \noalign{\hrule} height2pt&\omit&&\omit&&\omit&\cr } \hrule \halign{&\vrule#& \strut \quad\hfil#\hfil \quad\cr height2pt&\omit&&\omit&&\omit&\cr height2pt&\omit&&\omit&&\omit&\cr & $f/p$ && $\rho(\infty)$ && oscillation period $T_{{\rm osc}}$ && decay time $T$ & \cr height2pt&\omit&&\omit&&\omit&\cr \tablerule & $1 \cdot 10^{-2}$ && 0.376833 && $0.878 \pm 0.028$ && $0.988$ & \cr height2pt&\omit&&\omit&&\omit&\cr & $7 \cdot 10^{-3}$ && 0.381539 && $0.889 \pm 0.043$ && $1.155$ & \cr height2pt&\omit&&\omit&&\omit&\cr & $5 \cdot 10^{-3}$ && 0.385406 && $0.890 \pm 0.014$ && $1.312$ & \cr height2pt&\omit&&\omit&&\omit&\cr & $3 \cdot 10^{-3}$ && 0.390296 && $0.878 \pm 0.012$ && $1.588$ & \cr height2pt&\omit&&\omit&&\omit&\cr & $2 \cdot 10^{-3}$ && 0.393463 && $0.883 \pm 0.017$ && $1.719$ & \cr height2pt&\omit&&\omit&&\omit&\cr & $1.4\cdot10^{-3}$ && 0.395793 && $0.883 \pm 0.025$ && $1.966$ & \cr height2pt&\omit&&\omit&&\omit&\cr & $1 \cdot 10^{-3}$ && 0.397672 && $0.879 \pm 0.026$ && $2.296$ & \cr height2pt&\omit&&\omit&&\omit&\cr & $7 \cdot 10^{-4}$ && 0.399321 && $0.878 \pm 0.011$ && $2.455$ & \cr height2pt&\omit&&\omit&&\omit&\cr & $5 \cdot 10^{-4}$ && 0.400673 && $0.881 \pm 0.048$ && $2.603$ & \cr height2pt&\omit&&\omit&&\omit&\cr & $3 \cdot 10^{-4}$ && 0.402291 && $0.875 \pm 0.015$ && $3.645$ & \cr height2pt&\omit&&\omit&&\omit&\cr & $2 \cdot 10^{-4}$ && 0.403231 && $0.876 \pm 0.026$ && $3.559$ & \cr height2pt&\omit&&\omit&&\omit&\cr & $1 \cdot 10^{-4}$ && 0.404696 && $0.868 \pm 0.031$ && $3.951$ & \cr height2pt&\omit&&\omit&&\omit&\cr } \hrule}\hskip 1pt \vrule} \hbox{\quad \hbox{Table 1:} Estimates with $L=16384$ for the stationary density of trees $\rho(\infty)$ and} \hbox{\quad \phantom{Table 1:} the lowest decay mode in $\rho(t)$ of the two-dimensional forest-fire model.}} } \medskip\noindent These values are in good agreement with $$\rho_c - \rho(\infty) \sim \left({f \over p}\right)^{1/\delta} \, , \label{FORMrho}$$ where the critical density $\rho_c$ and the critical exponent $\delta$ are given by $$\eqalignno{ \rho_c &= 0.40844 \pm 0.00011 \, , &\eqnlabel{RESrhoC} \cr 1/\delta &= 0.466 \pm 0.004 \, . &\eqnlabel{RESdelta} }$$ These results agree within error bounds with the values obtained in \cite{14}, but our bounds are smaller. In particular we can now rule out that $1/\delta = 1/2$ as proposed by \cite{13}. \medskip\noindent Having determined $\rho(\infty)$, it is straightforward to extract the oscillation period and decay time of the slowest decay mode from $\rho(t)$. First, one determines those $t$ where $\rho(t)$ crosses the value $\rho(\infty)$. From the distances between these crossings the oscillation period $T_{{\rm osc}}$ can be determined easily. Averaging 10 to 15 half oscillation periods estimated in this manner for a suitable interval of time in Fig.\ 3 leads to the values given in Table 1. One observes that the oscillation period equals $T_{{\rm osc}} = 0.88$ within error bounds for all $f/p$, i.e.\ the slowest relaxational mode oscillates with a constant frequency. This is to be contrasted with the one-dimensional case where simulations of $\rho(t)$ clearly demonstrated that the oscillation period depends on $f/p$ \cite{15}. \medskip\noindent Finally, we extract the leading decay time $T$ from $\rho(t)$ according to the following procedure. At times $t$ precisely in the middle between two subsequent crossings used for the determination of the oscillation period, the value of $\abs{\rho(t) - \rho(\infty)}$ is determined. One finds for these (approximately ten) values of $t$ that $\abs{\rho(t) - \rho(\infty)} \sim \exp(-t/T)$ from which it is straightforward to obtain the estimates for $T$ presented in Table 1 \footnote{${}^{4})$}{ These estimates also verify that our equilibration times are long enough, since we have equilibrated the system for at least $6 T$ before starting to collect data. So, the non-stationary modes are damped by factors of at least $\exp(-6) \approx 2 \cdot 10^{-3}$ during data collection. }. These values for $T$ are compatible with a critical behaviour $$T \sim \left({f \over p}\right)^{-\zeta} \, , \label{DEFzeta}$$ where the critical exponent $\zeta$ is determined to be $$\zeta = 0.314 \pm 0.013 \, . \label{RESzeta}$$ This exponent probably is another new exponent that is not related to the ones determined so far \cite{12,14}. \bigskip\noindent \subsection{Summary of simulations} \medskip\noindent Table 2 summarizes our results for the critical exponents of the two-dimensional forest-fire model. It also includes the critical density $\rho_c$ and the global oscillation period which seems to be independent of $f/p$. For comparison we have also included the results for one dimension \cite{15,22}. In that case, the oscillation period diverges and we have listed the exponent rather than a period in Table 2. Similarly, the amplitude $a$ vanishes in one dimension for $f/p \to 0$ but is roughly constant in two dimensions. \medskip\noindent \centerline{\vbox{ \hbox{ \vrule \hskip 1pt \vbox{ \offinterlineskip \defheight2pt&\omit&&\omit&&\omit&\cr{height2pt&\omit&&\omit&&\omit&\cr} \def\tablerule{ height2pt&\omit&&\omit&&\omit&\cr \noalign{\hrule} height2pt&\omit&&\omit&&\omit&\cr } \hrule \halign{&\vrule#& \strut \quad\hfil#\hfil \quad\cr height2pt&\omit&&\omit&&\omit&\cr height2pt&\omit&&\omit&&\omit&\cr & quantity && value in $d=2$ && value in $d=1$ & \cr height2pt&\omit&&\omit&&\omit&\cr \tablerule & $\nu_T$ && $0.541 \pm 0.004$ && $0.8336 \pm 0.0036$ & \cr height2pt&\omit&&\omit&&\omit&\cr & $\eta_{{\rm occ}}$ && $0$ && $0$ & \cr height2pt&\omit&&\omit&&\omit&\cr & $\nu$ && $0.576 \pm 0.003$ && $1$ & \cr height2pt&\omit&&\omit&&\omit&\cr & $\eta$ && $0.36 \pm 0.03$ && $0$ & \cr height2pt&\omit&&\omit&&\omit&\cr & $\tau$ && $2.159 \pm 0.006$ && $2$ & \cr height2pt&\omit&&\omit&&\omit&\cr & $\lambda$ && $1.17 \pm 0.02$ && $1$ & \cr height2pt&\omit&&\omit&&\omit&\cr & $1/\delta$ && $0.466 \pm 0.004$ && $0$ & \cr height2pt&\omit&&\omit&&\omit&\cr & $\zeta$ && $0.314 \pm 0.013$ && $\approx 0.405$ & \cr \tablerule & $T_{{\rm osc}}$ && period: $0.88 \pm 0.02$ && exponent $\approx 0.194$ & \cr height2pt&\omit&&\omit&&\omit&\cr & $a$ && $0.030 \pm 0.001$ && $\sim (f/p)^{0.1031 \pm 0.0022}$ & \cr height2pt&\omit&&\omit&&\omit&\cr & $\rho_c$ && $0.40844 \pm 0.00011$ && $1$ & \cr height2pt&\omit&&\omit&&\omit&\cr } \hrule}\hskip 1pt \vrule} \hbox{\quad \hbox{Table 2:} Summary of our results for the critical behaviour of the} \hbox{\quad \phantom{Table 2:} forest-fire model.}} } \vfill \eject \section{Global models} \medskip\noindent We now wish to investigate through simplified variants of the model to what extent one can describe the stationary properties of the two-dimensional forest-fire model by global variables in a way similar to \cite{15}. There it was shown that in order to describe the one-dimensional critical stationary state it suffices to know the relative weight of the sum of all configurations with a fixed number of occupied (or empty) sites. Thus one is dealing with a kind of grand canonical ideal lattice gas. Such a model has no intrinsic spatial structure and leads to a two-point function independent of the distance. It can therefore be used to describe the asymptotic behaviour of the critical $C(y)$ of Section 2. Working with the (global) density of trees $\rho$, one has to specify the probability distribution $p(\rho)$ and obtains $C(y) = \langle \rho^2 \rangle - \langle \rho \rangle^2$ for $y \ne 0$ in the thermodynamic limit. This is positive for all continuous distributions. \medskip\noindent In two dimensions the limits $f/p \to 0$ and $L \to \infty$ do not commute with each other (in contrast to one dimension). Therefore, perturbation theory cannot be used to compute $p(\rho)$, and in fact we do not know of a good analytic method to determine $p(\rho)$ from the rules described in the Introduction. Therefore we use heuristic arguments and simulations to discuss it. \medskip\noindent The forest-fire model reminds one of site percolation. Thus, it is natural to try to relate the forest-fire model to percolation and gain some insight from that. Attempts in this direction have been made e.g.\ in \cite{8,21}. We will also try to use some results of percolation theory, but we will follow a different route. Namely, we try to interpret the stationary state of the forest-fire model at the critical point as a suitable ensemble of percolation problems with a distribution $p(\rho)$. \medskip\noindent It is known from percolation theory \cite{23} that in a homogeneous configuration with a density $\rho$ above the percolation threshold $\rho_{{\rm perc}}$ two arbitrary sites are connected with a finite probability. In such configurations, the dynamics of the forest-fire model with arbitrarily small $f/p > 0$ would very quickly destroy all percolating clusters and thus drive the density below the percolation threshold \footnote{${}^{5})$}{ This implies that the density $\rho$ cannot be continuous at $f/p=0$ for $d>1$ and is the reason why perturbation theory cannot be used in higher dimensions. }. Thus, the probability $p(\rho)$ to have a global density above the percolation threshold $\rho_{{\rm perc}}$ must vanish in the forest-fire model, i.e.\ $p(\rho) = 0$ for $\rho > \rho_{{\rm perc}}$. In one dimension one has $\rho_{{\rm perc}} = 1$ and in two dimensions $\rho_{{\rm perc}} = 0.592746$ \cite{23}. In all simulations $\rho(t)$ was well below this percolation threshold at all times (compare Fig.\ 3). \medskip\noindent It is useful to visualize the stationary state of the forest-fire model in order to gain some intuition. Fig.\ 4 shows an area of $760 \times 472$ sites at $t=66$ during a simulation with $L=16384$ and $f/p = 10^{-4}$ (compare also Fig.\ 2 of \cite{24}, Fig.\ 1 of \cite{14} and Fig.\ 6 of \cite{13}). One observes that at a certain fixed time the system consists of rather well-defined patches with different mean density of trees. These patches are not to be confused with single forest clusters, they usually contain many such clusters. Their typical size increases as $f/p$ becomes smaller, which reflects the divergence of correlation lengths. Looking at the time evolution of such a state \footnote{${}^{6})$}{ On X11 platforms, such a visualization is possible with the code used for the simulations presented in this paper. This code is available on the WWW \cite{19}.}, one observes an increase of density due to growth of trees that is constant throughout the patches and that lightning strikes essentially only the patches with the highest density. After a lightning has struck such a patch, a new patch with a low density is created. This new density is not really zero because there are always some trees in the patch that are not connected to the cluster which is destroyed. The idea now is that the important information about the critical stationary state is given by the distribution $p(\rho)$ of densities in these patches and that nothing essential changes if we replace such patchy systems with an ensemble of systems of {\it global} density $\rho$ occurring with the same probability $p(\rho)$. Of course, it remains to be tested to what extent this picture works. \medskip\noindent \centerline{\psfig{figure=patch.ps}} \smallskip\noindent {\par\noindent\leftskip=4.5 true pc \rightskip=4 true pc {\bf Fig.\ 4:} Snapshot of an area with $760 \times 472$ sites in the stationary state during a simulation with $L=16384$ and $f/p = 10^{-4}$. Trees are black and empty places white. \par\noindent} \bigskip\noindent \subsection{A simple model} \medskip\noindent Let us make a very simple model based on these ideas. Assume that below the percolation threshold $\rho_{{\rm perc}}$ trees just grow (with probability $p=1$) and no lightning strikes. As soon as the percolation threshold is exceeded, lightning strikes immediately and destroys all trees in the system (not just those in the percolating cluster). In order to determine $p(\rho)$ we first compute the mean lifetime of a configuration with density $\rho$, assuming that no lightning strikes (this consideration will also be useful later on). Such a configuration lives precisely $n$ local updates if $n-1$ times an occupied place and then an empty place are selected. Because of the above assumptions there is no correlation and therefore the probability of this to happen is given by $\rho^{n-1} (1-\rho)$. This yields the expectation value $t(\rho)$ of the lifetime as $$t(\rho) = \sum_{n=1}^{\infty} n \rho^{n-1} (1-\rho) = {1 \over 1-\rho} \, . \label{timeDens}$$ If no lightning strikes, the probability $p(\rho)$ to find a configuration with density $\rho$ is proportional to the lifetime of a state with this density. Taking into account that lightning strikes all trees at $\rho_{{\rm perc}}$, we find for this simple model $$p(\rho) = \cases{{\cal N} (1-\rho)^{-1}\, , & $\rho < \rho_{{\rm perc}}$, \cr 0 \, , & $\rho > \rho_{{\rm perc}}$, \cr} \label{simpleGmodel}$$ with the normalization constant given by ${\cal N}^{-1} = \int_0^{\rho_{{\rm perc}}} {\rm d}\rho \; (1-\rho)^{-1}$. In one dimension, \ref{simpleGmodel} agrees with eq.\ (5.4) of \cite{15} which was derived there using a different argument. \medskip\noindent This simple model has the following periodic time evolution: Trees grow until the density reaches the percolation threshold. Then lightning strikes and the process is restarted with a completely empty system. The average time needed for such a cycle is precisely the global oscillation time and is given by $\int_0^{\rho_{{\rm perc}}} {\rm d}\rho \; t(\rho)$. Inserting $t(\rho)$ according to \ref{timeDens} and the value of $\rho_{{\rm perc}}$ in two dimensions yields an oscillation period of $T_{{\rm osc}} \approx 0.898$ which is in very good agreement with what we found in simulations (compare Table 1). The mean density is given by $\int_0^1 {\rm d}\rho \; \rho p(\rho)$ from which one finds a critical density $\rho_c = 0.340\ldots$ (this deviates notably from the result \ref{RESrhoC} found by simulations of the full model). Finally, the probability to find two trees at arbitrary places is given by the second moment of $p(\rho)$, i.e.\ by $\int_0^1 {\rm d}\rho \; \rho^2 p(\rho)$. So, the two-point function exceeds the value $\rho_c^2$ by an amount $a=0.0289\ldots$ which agrees within error bounds with what we found by simulations for the large-distance asymptotics of the two-point function. Cluster-type quantities are not accessible as easily and would e.g.\ require again Monte-Carlo simulations. \medskip\noindent Although this simple model yields very good values for two quantities, its failure to give the correct $\rho_c$ is not surprising. Firstly, we have neglected the fact that lightning can already strike configurations with $\rho < \rho_{{\rm perc}}$ even for arbitrarily small $f/p > 0$ (compare Appendix A). Secondly, lightning does not really lead to the completely empty system, but usually leaves some isolated trees or small clusters in the patch behind. Unfortunately we do not know how to treat either effect analytically. This lack of knowledge also has the effect that we can in general not compute an oscillation time from $p(\rho)$ although we do of course still think of the system as evolving in cycles (compare also section 6.3.2 of \cite{25}). \bigskip\noindent \subsection{Realistic distributions} \medskip\noindent Next we try to obtain a realistic $p(\rho)$ from Monte-Carlo simulations. Measurements of the global density cannot be used to extract $p(\rho)$ from a simulation, because $\rho$ fluctuates only very little around its mean value for sufficiently large systems (see Fig.\ 3). Therefore, one has to look at the distribution of local densities. In a large system, areas with different local densities coexist and we are interested precisely in these local fluctuations and not just the global average. We have decided to divide the system into $16 \times 16$ plaquettes and use the distribution of the average density per plaquette. This size of the plaquettes was chosen because then their linear extent is much smaller than the correlation lengths and they contain sufficiently many sites to obtain a fairly smooth distribution. Fig.\ 5 shows a result $p_{{\rm re}}(\rho)$ obtained in this manner using the parameter values closest to the critical point, namely $f/p = 10^{-4}$ and $L=16384$. Samples were taken at the same 90 times where also the correlation functions were determined, amounting to a total of almost $10^{8}$ samples for local densities. The normalization in Fig.\ 5 is such that $\sum_{r=0}^{256} p_{{\rm re}}(r/256) = 1$. \medskip\noindent As explained above, the first and second moments of $p(\rho)$ are related to the mean density and the asymptotic value of the two-point function. From this one finds $$\rho(\infty) = 0.4044\ldots \, , \qquad a = 0.031\ldots \, . \label{ParamRealModel}$$ The way we have determined $p_{{\rm re}}(\rho)$ ensures that the first moment indeed equals the value obtained by directly taking a global average. The slight difference between \ref{ParamRealModel} and the value for $\rho(\infty)$ in Table 1 is due to the fact that the latter is based on a much larger amount of configurations. As for the simple model presented before, the prediction for $C(y) = a \approx 0.031$ agrees within error bounds with what we expect for the asymptotics of the two-point function at the critical point. \medskip\noindent \centerline{\psfig{figure=dens_fluc.ps}} \smallskip\noindent {\par\noindent\leftskip=4.5 true pc \rightskip=4 true pc {\bf Fig.\ 5:} Result for the distribution $p_{{\rm re}}(\rho)$ on $16 \times 16$ plaquettes in a simulation with $L=16384$ and $f/p = 10^{-4}$. \par\noindent} \medskip\noindent Even though this basic test yields good values, one should be aware that examining the system only in windows blurs the distribution in Fig.\ 5 in several ways. Firstly, looking through a window containing only 256 sites, one finds a smearing of the density by $\Delta \rho \approx 0.02$ just because of statistical effects. Secondly, such windows may accidentally intersect the boundary between two patches with low and high densities. This has the effect that $p(\rho)$ is estimated too large for intermediate $\rho$ and too small for the extremal ones. Our choice of $16 \times 16$ plaquettes is designed to make both effects reasonably small and is about the best we can do. \medskip\noindent Let us now discuss Fig.\ 5 keeping these effects in mind. For not too small values of $\rho$, $p_{{\rm re}}(\rho)$ increases slowly and reaches a maximum around $\rho \approx 0.54$. Around $\rho \approx 0.62$ there is a sharp decrease. The broad distribution shows that fluctuations of $\rho$ are important. A peak just below the percolation threshold $\rho_{{\rm perc}}$ and a steep decrease above it correspond to our expectation. However, there is still a substantial contribution to $p_{{\rm re}}(\rho)$ above $\rho_{{\rm perc}}$ which is not explained by windowing effects. This is due to the patchy structure of the system: Finite patches with $\rho > \rho_{{\rm perc}}$ are not destroyed instantly, but rather live for a time which is the longer the smaller these patches actually are. One could also say that the patchy structure demonstrates that the system is actually correlated. In particular at small distances ($y \ale 20$), the two-point function $C(y)$ retains a $y$-depence (compare Section 2.1) for $f/p \to 0$ and thereby exceeds the asymptotic constant $a$. This correlation is necessary to observe a non-trivial distribution of local densities, but also leads to contributions to $p(\rho)$ above the percolation threshold. \medskip\noindent So, if we want to work with a `realistic' $p(\rho)$, the best we can do is to proceed with the one shown in Fig.\ 5. Alternatively, one can work with an approximation to this distribution where one suppresses the undesired $p(\rho)$ above the percolation threshold by hand. One such approximation we have examined is a linear one, i.e\ $p_{{\rm lin}}(\rho) \sim \rho$ for $\rho < 0.59$ and $p_{{\rm lin}}(\rho) = 0$ for $\rho > 0.59$. This linear distribution yields $\rho_c \approx 0.395$ and $a \approx 0.019$ (both somewhat too small). \medskip\noindent \centerline{\psfig{figure=ns_fluc.ps}} \smallskip\noindent {\par\noindent\leftskip=4.5 true pc \rightskip=4 true pc {\bf Fig.\ 6:} The cluster-size distribution obtained from a simulation of the full model with $L=16384$ and $f/p = 10^{-4}$ (full line), the one based on $p_{{\rm re}}(\rho)$ (long dashes) and the result obtained from $p_{{\rm lin}}(\rho)$ (short dashes). \par\noindent} \medskip\noindent The determination of the cluster distribution $n(s)$ and $K(y)$ is a complicated combinatorial problem which we solve again using simulations. One creates configurations with densities distributed according to the given $p(\rho)$ and then measures $n(s)$ and $K(y)$. We have created 100000 configurations on a $512 \times 512$ lattice distributed according to $p_{{\rm re}}(\rho)$ and 70000 configurations on a $1024 \times 1024$ lattice for $p_{{\rm lin}}(\rho)$. Fig.\ 6 shows the cluster-size distribution $n(s)$ obtained in this manner together with the one obtained from a simulation of the full model. For small cluster sizes ($s < 100$), all three distributions are close to each other. However, at larger $s$ the distributions based on a globally given $p(\rho)$ decay faster than the true $n(s)$. The corresponding exponent is $\tau \approx 2.48$ for the distribution $p_{{\rm re}}(\rho)$ and $\tau \approx 2.44$ for $p_{{\rm lin}}(\rho)$ -- both much closer to the mean-field value $\tau = 5/2$ \cite{21} than to the true value \ref{REStau}. In Fig.\ 6 one also observes a peak in the cluster-size distribution corresponding to $p_{{\rm re}}(\rho)$ for $1 \cdot 10^5 \le s \le 2.5 \cdot 10^5 \approx 512^2$, i.e.\ just below the volume of the system. This is due to the non-vanishing of $p_{{\rm re}}(\rho)$ for $\rho > \rho_{{\rm perc}}$ which leads to clusters spanning a finite (and large) fraction of the system. \medskip\noindent \centerline{\psfig{figure=cor_fluc.ps}} \smallskip\noindent {\par\noindent\leftskip=4.5 true pc \rightskip=4 true pc {\bf Fig.\ 7:} The full line shows the correlation function $K(y)$ obtained from a simulation of the full model with $L=16384$ and $f/p = 10^{-4}$. The result for $p_{{\rm re}}(\rho)$ is shown by the line with long dashes, the one obtained from $p_{{\rm lin}}(\rho)$ is indicated by the shorter dashes. \par\noindent} \medskip\noindent Fig.\ 7 shows the probability $K(y)$ to find two trees at distance $y$ inside the same cluster. Here, the results obtained from the two $p(\rho)$'s deviate more notably from the result obtained by simulation of the full model (full line). As noted above, the distribution $p_{{\rm re}}(\rho)$ gives rise to percolating clusters and produces a constant background in $K(y)$ of approximately $0.0552$. After subtracting this background, one can fit $K(y)$ with \ref{DEFeta} for $\xi_c \approx 180$ and $\eta \approx 0.92$. This value for $\eta$ is more than twice as large the true one \ref{RESeta}. For $p_{{\rm lin}}(\rho)$ one finds good agreement with the form \ref{DEFeta} for $\xi_c = L/2$ and $\eta \approx 0.95$. \bigskip\noindent \subsection{Summary and outlook on global models} \medskip\noindent We have shown that the distributions $p_{{\rm re}}$ and $p_{{\rm lin}}$ lead to a power law for $n(s)$. One can check that the same is true for other $p(\rho)$'s that are cut off at $\rho_{{\rm perc}}$ in a way similar to $p_{{\rm lin}}$. Thus, a power law in $n(s)$ arises automatically from a description in terms of global quantities and need not be a signal for criticality in a conventional sense. However, in all examples for $p(\rho)$ discussed so far, we obtained values for $\tau$ and $\eta$ that are unsatisfactorily larger than those of the full model. Nevertheless, a description in terms of $p(\rho)$ can still be forced to work because for two-dimensional critical percolation one has $\tau \approx 2.055$ and $\eta \approx 0.208$ \cite{23} -- values which are smaller than the ones in Table 2. So, one can obtain the desired value e.g.\ of $\tau$ by peaking the distribution $p(\rho)$ more prominently just below $\rho_{{\rm perc}}$. Adjusting just $\tau$ to its correct value can be expected to also give a reasonably good approximation for $\eta$. Afterwards, both $\rho_c$ and $a$ may be tuned to the desired values by adding another peak in $p(\rho)$ for smaller densities (which does not contribute to large clusters and thus affect the asymptotics of cluster quantities). However, there does not seem to be a natural way to make these adjustments. In particular, local densities above $\rho_{{\rm perc}}$ do exist in the full model which would have to be discarded by hand in a global model in order to obtain the correct $K(y)$. \medskip\noindent In one dimension the spatial structure becomes irrelevant at the critical point \cite{15}. We have seen in this Section that this can be generalized to the qualitative features of the critical correlations in two dimensions, but not to the quantitative details. In contrast to the one-dimensional case we had no analytical tools at our disposal and have therefore not been able to derive the distribution $p(\rho)$ of local densities explicitly. The introduction of block-spin variables is reminiscent of real-space renormalization group ideas and it would be interesting to see if they can be used to find $p(\rho)$. However, one would have to go beyond the block-spin renormalization-group study of \cite{18}. Firstly, one would have to admit densities different from 0 or 1 for the block-spin variables, and moreover the dynamics should not be treated just in mean-field approximation. \medskip\noindent Two-dimensional percolation is believed to be conformally invariant (see e.g.\ \cite{26}). The globalized models are just ensembles of percolation problems and should therefore be conformally invariant as well. It would be interesting to know if also the stationary state of the full model in two space dimensions is conformally invariant, even if the standard techniques of conformal field theory would probably not say much about quantities like cluster sizes. \bigskip\noindent \section{Conclusions} \medskip\noindent In this paper we have again looked at several aspects of the two-dimensional forest-fire model. Firstly, we have shown that the two length scales $\xi$ and $\xi_c$ have different critical exponents. That this might be possible had been suggested by a study of the one-dimensional model \cite{15} which illustrates that one-dimensional systems can provide useful insights because of their relative simplicity even if one is actually interested in higher-dimensional versions. This result shows that in generic non-equilibrium systems geometric objects and the usual (occupancy) correlation functions can behave completely differently. For an equilibrium system as the Ising model such an observation was already made some time ago in \cite{27}. In this case, percolation occurs away from the critical point, i.e.\ the two length scales are so different that they diverge at different temperatures (see e.g.\ \cite{28,23}). \medskip\noindent In order to show that $\nu_T \ne \nu$ we had to improve the error bounds of earlier investigations. As a by-product we have also improved the accuracy of other critical exponents. It may be possible that one could still improve the error bounds by another digit using optimized code on today's most powerful computers. Historically, Monte-Carlo simulations have already several times lead to values for the critical exponents that had to be corrected later on \cite{8,12,13,14} and as we have shown here, some of them were still not treated adequately. Therefore, it may be desirable to perform yet another independent verification of the results presented here, but a further increase of accuracy may not be necessary for this end. \medskip\noindent The second part of the paper focussed on a globalized model. We found that one can easily obtain power laws in the cluster-size distribution by discarding the spatial structure, i.e.\ by making the usual two-point function independent of the spatial coordinates. This generalizes a result obtained for one dimension in a previous paper \cite{15} and is line with the observation in \cite{16}, based on a different one-dimensional model, that one can obtain power laws in clusters or avalanches by global (`coherent') driving. However, some quantitative predictions of the globalized model did not work out satisfactorily. One reason is that the full two-dimensional forest-fire model has non-trivial two-point functions at least at small distances which is also reflected by the existence of patches with local densities above the percolation threshold. In addition, the full model exhibits many critical exponents that cannot be described by a global model. \medskip\noindent A study of the usual correlation functions would also be desirable in other models of self-organized criticality where they have not yet been investigated. This could help to clarify to what extent the process of self-organization can be regarded as a global phenomenon. Two-point correlation functions would also be important quantities to examine in experiments. It would e.g.\ be interesting to extract the spatial correlation functions of the local heights and slopes from the experimental data of \cite{3}. \medskip\noindent Even for the two-dimensional forest-fire model there are still many issues we have not looked at, including e.g.\ finite-size effects. We have only looked at the regime $f V \gg p$ which is close to the thermodynamic limit. One could also look at a different limit, namely $f V \ll p$ where the first-order approximation of \cite{15} applies independent of the spatial dimension. In this limit, trees grow until the lattice is full and after a certain time of rest, lightning destroys all these trees and the process starts again. It would be interesting to see how this behaviour crosses over to the critical behaviour studied here as the volume of the system is increased, and if finite-size scaling can be observed. \vskip 0.8 cm \displayhead{Acknowledgments} \medskip\noindent Useful discussions with S.\ Clar, M.\ Hasenbusch and D.\ Stauffer are gratefully acknowledged. The work of A.H.\ has been funded by the Deutsche Forschungsgemeinschaft. \sectionnumstyle{Alphabetic} \newsubsectionnum=\z@\@resetnum\sectionnum=0 \vskip 1.6 cm
proofpile-arXiv_065-451
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} In this work we will study the behaviour of the $\rm Q \cdot Q$ interaction in a nucleus as a function of the strength of the interaction. Although this is a model study it helps to make things concrete by focusing on a particular nucleus. We choose $^{10}Be$. This nucleus is of particular interest because in a $0p$ space calculation with $\rm Q \cdot Q$ there are some interesting degeneracies. For example, the $2_1^+$ and $2_2^+$ states are degenerate, and both have orbital symmetry [400]. The $L=1~S=1$ states with orbital symmetry [330] and [411] are degenerate. Thus we have two sets of degenerate triplets $J=0^+,~1^+$ and $2^+$ emanating from the above two orbital configurations. These degeneracies can easily be found by applying the $SU(3)$ formula: \[E(\lambda \mu)=\bar{\chi}\left[-4(\lambda^2+\mu^2+\lambda\mu+ 3 (\lambda+\mu))+3L(L+1) \right] \] If we write the interaction as $\chi Q \cdot Q$ then, in the $0p$ space and when we change $\chi$, the degeneracies will not be removed. All that will happen is that the energies of all states will be proportional to $\chi$, and so the spectrum will be blown up or shrunk as $\chi$ is made larger or smaller. What happens though if we include contributions from other major shells? The $\rm Q \cdot Q$ interaction will not connect $\Delta N=1$ states because of parity arguments, but there will be $\Delta N=2$ admixtures. Indeed, the concept of an $E2$ effective charge is often illustrated by using a $\rm Q \cdot Q$ interaction to obtain $\Delta N=2$ admixtures. \section{The Hamiltonian and Choice of Parameters} It has been shown by Bes and Sorensen~\cite{bes} that if in the valence space ($0p$) the appropriate $\rm Q \cdot Q$ interaction is $-\chi_0 Q \cdot Q$, then in a space which includes $\Delta N = 2$ excitations the appropriate strength is $-\frac{\chi}{2} Q \cdot Q$. We wish to vary the interaction, and we therefore parameterize it as \[V=-t \left (\frac{\chi_0}{2} \right ) Q \cdot Q \] \noindent so that for $t=1$ we have the standard choice of Bes and Sorensen~\cite{bes}. For $^{10}Be$, we have $\chi_0= 0.36146~MeVfm^{-1}$. The Hamiltonian we use is \[ H=\sum_i T_i + \frac{1}{2}m \omega^2 r_i^2 - \sum_{i < j} t \frac{\chi_0}{2} Q(i) \cdot Q(j) \] \noindent We perform shell model calculations using the OXBASH shell model code~\cite{oxbash} in the space $(0p)^6$ plus $2 \hbar \omega$ excitations. As mentioned in the introduction, when harmonic oscillator wave functions are used in a single major shell, the single particle terms in the above Hamiltonian are constant, so that the only part of the Hamiltonian that affects the spectra is the two-body term. The separation of energies is linear in $t$ -the wave functions are unaffected by the choice of (positive) $t$. However, when $\Delta N =2$ excitations are allowed, the linear terms are no longer constant, and the behaviour as a function of $t$ is more complicated. We will now study this behaviour as a function of $t$ in $^{10}Be$. \subsection{Removal of Degeneracies} In Table I we present $t=1$ results for $^{10}Be$ in a large space $(0p)^6$ plus all $2 \hbar \omega$ excitations. We present the results for the energies of $J=1^+$ and $2^+$ $T=1$ states, as well as $B(M1)$ and $B(E2)$ transitions from the ground state to these states. More precisely, it is the isovector $B(M1)$ in units of $\mu_N^2$ and isoscalar $B(E2)$ ($e_p=1,~e_n=1$) in units of $e^2 fm^4$. We see that there are many degeneracies still present in the large space $e.g.$ four $J=1^+~T=1$ states at 12.12 $MeV$ and three at 13.90 $MeV$, as well as three $J=2^+~T=1$ states at 12.12 $MeV$. These degeneracies clearly correspond to various $S,~T$ combinations for states of given $L$ and orbital symmetry [$f$]. However, other `accidental' degeneracies which were present in the small space $(0p)^6$ are no longer present. The $J=1^+_1$ and $1^+_2$ states at 3.74 $MeV$ and 7.31 $MeV$ are linear combinations of the $L=1$ [330] and $L=1$ [411] configurations (actually they are parts of $J=0^+,~1^+,~2^+$ triplets). In the small space, these two $J=1^+$ states (or triplets) were degenerate -now one of the states is almost at twice the excitation energy of the other. The $2_1^+$ and $2_2^+$ states are no longer degenerate. The $2_1^+$ state is at 2.19 $MeV$ with $B(E2)\uparrow$ from the ground state of 63.8 $e^2 fm^4$, whilst the $2_2^+$ state is at 3.40 $MeV$ with $B(E2)\uparrow=113.4~e^2 fm^4$. Note that, contrary to experiment, the second $2^+$ state is the one most strongly excited. When a reasonable spin-orbit interaction is added to the Hamiltonian the situation is corrected. It should be pointed out that in perturbation theory, in which only the {\em direct} part of the particle-hole interaction of $\rm Q \cdot Q$ is used to renormalize the interaction between two particles in the valence space, the degeneracies above would {\em not} be removed. The relevant diagram is the familiar Bertsch-Kuo-Brown bubble (or phonon) exchange between two nucleons~\cite{bert,kbrown}. For a simple $\rm Q \cdot Q$ interaction, this diagram simply renormalizes the strength of the $\rm Q \cdot Q$ interaction. Clearly, changing the strength in the valence space will not remove the degeneracies. Thus, the shell model diagonalization implicitly contains effects beyond the direct bubble diagram. Furthermore, these effects are quite important. \subsection{Change in the Nature of the Ground State as $t$ Increases} We now vary $t$ over the range $0 < t < 2$. In Fig. 1 we plot as a dot-dash curve the value of $E/t$ for the lowest $2^+$ state, the one with finite but small $B(E2)$ strength from ground. We also plot $E/t$ as a solid line for the state with the strongest $B(E2)$ from ground. It starts off at $t=1$ as the second $2^+$ state. In the $0p$ space, the $2_1^+$ and $2_2^+$ states would be degenerate, and the curve for $E/t$ vs. $t$ would be a horizontal line. However, in the $0p+2\hbar\omega$ space there is a dependence on $t$ (and more so for the solid curve). But, for $t\approx 1.8$ and beyond, all the $B(E2)$ strength from ground state goes to the new lowest $2^+$ state. Furthermore, the value of $E/t$ becomes constant for $t \geq 1.8$ $i.e.$ the curve becomes horizontal. Clearly, the nature of the ground state changed beyond $t=1.7$. We will now examine this change in more detail. If the only thing that happened was that there was a new $J=0^+$ ground state, then of course there would be a sudden change in the $B(E2)\uparrow$'s from this new ground state to the $2^+$ states. However, the static quadrupole moments of the $2^+$ states themselves would not change. For $t=1.1$ the $B(E2)\uparrow$ to the $2_1^+$ state is 40.61 $e^2 fm^4$ and to the $2^+_2$ state 158.0 $e^2 fm^4$. The calculated static quadrupole moments are respectively 11.56 $e fm^2$ and -11.46 $e fm^2$. As discussed by Fayache, Sharma and Zamick in Ref.~\cite{qqt}, this is consistent with the two $2^+$ states being prolate, with about the same intrinsic quadrupole moment $\rm Q_0$, but the lower state would have $K=2$ and the upper one $K=0$. Indeed, for $K=0$ $\rm Q(2^+)=-2/7Q_0$, and for $K=2$ $\rm Q(2^+)=+2/7Q_0$. The behaviour for $t=1.1$ is maintained for $t=1.3$ and 1.5, but for $t=1.7$ and up to $t=1.75$ there is a big drop in $B(E2)_{0_1^+ \rightarrow 2^+_2}$. The other three quantities do not change, not even -strangely enough- $\rm Q(2^+_2)$. Remember that for $t=1.75$ we are still below the critical $t$ for which the $J=0^+$ ground state changes its nature. What is clearly happening is that a third $2^+$ state has crossed over and came below what was formerly the $2^+_2$ state. This is confirmed by noting that at $t=1.75$ the $B(E2)_{0^+_1 \rightarrow 2^+_3}$ is very large (162.1 $e^2 fm^4$). Clearly, in going from $t=1.7$ to $t=1.75$, the $2^+_2$ and $2^+_3$ states have interchanged positions. The fact that the static quadrupole moments of the two states are about the same means that both states can be associated with two different prolate $K=0$ bands which have the same deformation. Next we consider $t=1.76$. Here we are just beyond the critical $t$, and the nature of the ground state has changed. Now the $B(E2)$ to the $2^+_1$ state is much weaker, and the $B(E2)$ to the $2^+_2$ state is strong. This suggests that the new $0^+$ ground state and the new $2^+_2$ state are members of a new rotational band, and that both of these states have come down in energy together. When we go from $t=1.76$ to $t=1.78$ there is a big change, but the results stabilize beyond that. Now the $B(E2)$ to the $2_1^+$ state is the strongest (140 $e^2 fm^4$), and the {\em signs} of the static quadrupole moments change. Now $\rm Q(2^+_1)$ is negative, and $\rm Q(2^+_2)$ is positive. What is clearly happening is that there is another cross-over. What was formerly the $2^+_3$ state at $t=1.7$ first crosses the the $2^+_2$ state at $t=1.75$ (as mentioned above), and now crosses the $2^+_1$ state at $t=1.78$. By $t=1.78$ and beyond, the $0^+$ and $2^+$ members of a new band have become the lowest two states, and the results stabilize. \section{Interpretation of the New Band: States with Integer Occupancies} In Fig. 2 we plot the rapid descent of the $J=0^+$ and $2^+$ members of the new rotational band. We start from $t=1$, but if we project backward we see that for small $t$ the band emanates from the $2 \hbar \omega$ region. To better ascertain the nature of the new band, we give in Table III the occupancies of the single-particle levels that were used in this calculation {\em i.e.} $0s$, $0p$, $1s-0d$ and $1p-0f$. At $t=1.7$, just before the critical value, the $J=0^+$ ground state is normal. The occupancy of the four major shells (in the order mentioned above) is 3.84, 5.78, 0.18 and 0.20. The first excited $0^+$ state at 0.585 $MeV$ has occupancy 4, 4, 2 and 0. Clearly two nucleons have been excited from $0p$ to $1s-0d$. What is at first surprising is that, for this state, the occupancies are {\em precisely integers}. This is not an isolated example. It is also true at $t=1.7$ for the second $2^+$ state at 4.33 $MeV$. When we go to $t=1.9$, we have passed the critical value, and things have settled down. the lowest $0^+$ and $2^+$ states are now the $2p-2h$ states, both with the integer occupancies 4, 4, 2 and 0. Whereas most states do not have integer occupancies, there are many which do. These are at higher energies. For example, for $t=1.7$, there are other states with the occupancy 4, 4, 2, 0 at 21.64, 22.95, 22.97, 25.32 and 35.75 $MeV$. These are states with occupancy (3, 4, 1, 0) at 43.6 and 44.2 $MeV$. The latter correspond to lifting one nucleon through two major shells. Why do we get such a simple behaviour for the $2p-2h$ states? The answer involves a special feature of the $\rm Q \cdot Q$ interaction: all matrix elements in which two particles in a major shell $N$ scatter into a major shell $N \pm 1$ vanish. This is due to a parity selection rule. For example, $\langle 0p~0p | Q\cdot Q | 0d~0d \rangle$ factors into $\langle 0p |Q | 0d \rangle \langle 0p |Q | 0d \rangle$, and each of these factors vanishes because of this parity rule. Carrying the argument further, there can be no matrix element coupling the $(0s)^4(0p)^4(0d-1s)^2$ configuration with other configurations such as $(0s)^4(0p)^6$ or $(0s)^4(0p)^5(0f-1p)$ etc... Also, in our calculation we have limited the space to 2 $\hbar \omega$ excitations. Once we create the state with occupancy (4,4,2,0) our model space does not permit further excitations. This explains the integer occupancy. Presumably if we enlarged the space to include 4 $\hbar \omega$ excitations we would no longer have the integer occupancies. It would also be of interest to study the $4p-4h$ states. This also explains why, as we increase $t$ towards 1.76, the descent of the new band is so simple. Since there is no mixing with the other configurations, the $2p-2h$ $J=0^+$ and $2^+$ states can just slip down below the $(0p)^6$ states. The rapid descent of the $2p-2h$ states can be understood in terms of the Nilsson model. To form the $2p-2h$ state, we take two nucleons from the $0p$ shell and put them in the Nilsson orbit ($Nm_3\Lambda$) with quantum numbers (220). This orbit comes down rapidly in energy as the nuclear deformation is increased. The Nilsson one-body deformed Hamiltonian can be obtained from the $\rm Q \cdot Q$ two-body interaction by replacing $\rm Q \cdot Q$ by $\rm Q \cdot \langle Q \rangle$ where $\langle Q \rangle$ is the quadrupole moment of the intrinsic state. \section{Closing Remarks} In this work, we have studied the properties of the interaction $-t\frac{\chi_0}{2} Q \cdot Q$ as a function of the coupling strength $t$ in an extended model space which includes all $2 \hbar \omega$ excitations beyond the valence space. Using $^{10}Be$ as an example, we found that states that were `accidentally' degenerate in the $0p$ valence space ($e.g.$ $2_1^+$ and $2_2^+$ of orbital symmetry [42], or the $J=0^+,1^+,2^+$ triplets of orbital symmetries [411] and [33]), are no longer degenerate in the extended space. This means that the $2 \hbar \omega$ admixtures do more than renormalize the coupling strength of $\rm Q \cdot Q$. The extended model space allows for $2p-2h$ admixtures and indeed, for sufficiently large $t$ ($t \geq 1.8$), the $J=0^+$ and $2^+$ members of this new band become the new $0_1^+$ and $2_1^+$ states. We find that this band, unlike the `normal' ground state band for $t < 1.7$, has integer occupancies 4, 4, 2 and 0 for $0s$, $0p$, $1s-0d$ and $1f-0p$ respectively. There is no mixing between the new $2p-2h$ band and the $0p-0h$ band. This is a special feature of the $\rm Q \cdot Q$ interaction. On a speculative level we may argue that, for the $2p-2h$ band above, we should use a value of $t$ considerably larger than 1. We don't have enough model space to renormalize the two-body interaction in this band. Just as the interaction between the $(0p)^6$ states gets renormalized by the configurations in which one nucleon is excited through 2 major shells, so the interaction for the $2p-2h$ state would get renormalized by allowing at least one nucleon to be excited through 2 major shells. But this would be a 4 $\hbar \omega$ state which, for practical reasons, we don't have in our model space. We can use the $\rm Q \cdot Q$ interaction to place these $2p-2h$ bands at the correct energies, but we will need other components of the realistic nucleon-nucleon interaction in order to mix these bands with the $0p-0h$ bands. For example, we could use the dipole-dipole or octupole-octupole parts of the nucleon-nucleon interaction. Alternatively one can work directly with realistic interactions. There has been much progress in large-basis shell model calculations in light nuclei. For example, there is the work of W.C. Haxton and C. Johnson~\cite{hax} where they actually get the superdeformed $4p-4h$ state in $^{16}O$ at a reasonable energy, although perhaps not with the full quadrupole collectivity. There is also the work of Zheng $et. al.$~\cite{zheng} where up to $8 \hbar \omega$ excitations have been included in calculations of nuclei ranging from $^4He$ to $^7Li$. Also, Zamick, Zheng and Fayache~\cite{zzf} required multi-shell admixtures to demonstrate the `self-weakening mechanism' of the tensor interaction in nuclei. Nevertheless, schematic interactions like $\rm Q \cdot Q$ still play a primary role in describing nuclear collectivity throughout the periodic table. They are of special importance for highly deformed intruder states. Surprisingly, there have been very few studies of schematic interactions in multi-shell spaces. In the Elliott $SU(3)$ model, momentum terms have been introduced to prevent $N=2$ admixtures in the valence space~\cite{elliott}. This has lead to great simplicities and beautiful results. There have been $R.P.A.$ studies with $\rm Q \cdot Q$ which involve $N=2$ mixing. These studies give $E2$ effective charge renormalizations in the valence space and also the energies of the giant quadrupole resonances, but they will not give us the highly deformed states such as the $2p-2h$ state that we have found here. We therefore feel that careful studies of the schematic interactions in multi-shell spaces are important, and we hope that others will agree. \acknowledgements This work was supported by a Department of Energy grant DE-FG02-95ER 40940. One of us (L.Z.) thanks E. Vogt, B. Jennings, H. Fearing, and P.Jackson for their help and hospitality at TRIUMF.
proofpile-arXiv_065-452
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\subsection{Light yield measurement} The absolute light yield for minimum--ionising particles was measured in a cosmic ray telescope with an effective area of 12~x~12 cm$^{2}$. The PMT pulse was integrated by a LeCroy 2249A ADC, triggered by a threefold coincidence of the signals from the cosmic trigger\footnote{The absolute trigger efficiency was determined to be 98.9\%}. The ADC gate length was 100 ns. Figure<~\ref{cos.ps} shows the response to cosmic ray particles for the final tile/fiber/PMT layout. \begin{figure}[tbp] \centerline{\psfig{figure=DESY-96-139_1.ps,bbllx=0pt,bblly=140pt,bburx=550pt,bbury=650pt,height=8cm}} \caption{\it Response of a scintillator tile to cosmic muons. The peak around the pedestal value of 45 ADC counts is due to false triggers and zero response due to photostatistics.} \label{cos.ps} \end{figure} The ratio of triggers with no signal above pedestal recorded in the ADC to the total number of triggers is equal to (0.37$\pm$0.03\%), taking into account the trigger efficiency. From this we obtain an absolute light yield of 5.6$\pm$.1 photoelectrons. \subsection{Uniformity measurement} The tile uniformity was measured with a collimated $^{106}$Ru source, a computer-controlled x-y scanning table as scintillator support and an XP2020 photomultiplier. The ADC was triggered by a coincidence of two photomultiplier signals reading out a large trigger counter located underneath the tile to be tested (figure~\ref{tab.ps}). To ensure that the $^{106}$Ru source simulates minimum--ionising particles, we measured the light yield of the tile also with the cosmic telescope. The pulse height spectra of cosmic muons and electrons from the $^{106}$Ru source are very similar in shape, with peak values equal within 5\%. A 3--mm--diameter collimator was used to measure the response over a 20~x~20 cm$^{2}$ tile in steps of 2~mm in x and y. The response is uniform within 5\%. The distribution of the mean values of the measurements for all positions is shown in figure~\ref{mean.ps}. The slight bump below 120 channels is an effect occurring near the edge of the tile where some particles leave the tile before crossing its full thickness. \begin{figure}[tbp] \centerline{\psfig{figure=DESY-96-139_2.ps,bbllx=130pt,bblly=55pt,bburx=450pt,bbury=600pt,height=8cm,angle=270}} \caption{\it Scanning table uniformity measurement} \label{tab.ps} \end{figure} \begin{figure}[tbp] \centerline{\psfig{figure=DESY-96-139_3.ps,bbllx=15pt,bblly=150pt,bburx=550pt,bbury=670pt,height=8cm}} \caption{\it Spread of response over a tile. The response is uniform within 5\%.} \label{mean.ps} \end{figure} \section {Mechanical layout} \subsection {Tile assembly} The scintillator tiles are assembled in cassettes, made of 0.4~mm thick stainless steel; they are 20~cm wide and vary in length, containing between 1 and 10 tiles. In order for the readout fibers to clear the neighboring scintillator tile the 5~mm thick tiles are raised at the readout end, supported by a 2.5~mm thick rohacell strip. The total thickness of the cassette is $\simeq$ 11~mm and represents about 5~\% of a radiation length (1.2~\% for the scintillator and 4~\% for the stainless steel). The length of readout fibers is $\simeq$ 300~cm. The six readout fibers of one tile are glued together in a connector which is fixed to the PMT housing. In addition to the six fibers a seventh clear fiber is added to guide the light from a laser/LED system to the PMT. \subsection {Detector assembly} The forward and rear calorimeters are split in two halves, such that they can be withdrawn from the beam pipe region during injection of the electrons and protons in HERA. A group of 19~cassettes, which covers one half of each calorimeter face, is glued on a 2~mm thick aluminium plate (0.02 radiation lengths) of 2~x~4~m$^2$. Figure~\ref{layout_presam} shows the coverage of the calorimeter by the presampler. Shown is the segmentation of the electromagnetic sections, which is finer in the region not shadowed from the nominal interaction point by the barrel calorimeter. The 20~x~20~cm$^{2}$ towers covered by the presampler tiles are shaded. \begin{figure}[tbp] \centerline{ \psfig{figure=DESY-96-139_4.ps,bbllx=0pt,bblly=100pt,bburx=580pt,bbury=680pt,height=8cm} \psfig{figure=DESY-96-139_5.ps,bbllx=0pt,bblly=100pt,bburx=580pt,bbury=680pt,height=8cm}} \caption{\it Front view of the forward (FCAL) and rear calorimeter (RCAL). The 20~x~20~cm$^{2}$ white square in the center corresponds to the hole for the beam pipe. The coverage of the presampler is indicated by the shaded region.} \label{layout_presam} \end{figure} A 2.5~mm diameter tube is glued over the full length on the outside of each cassette, positioned at the center of the tiles. The tube guides a radioactive source for calibrating the light output of the individual tiles and the gain of the PMT channels (see section 7.2). \section {Photomultiplier tests} \subsection{Performance specifications} Due to the limited space available in the ZEUS detector it was decided to use multi\-chan\-nel PMT's. The magnetic field amounts to a few hundred Gauss in the area where the PMT's are located and therefore adequate shielding is needed. Since we measure pulse heights, the crosstalk between adjacent channels in the tube is required to be less than 5\%. Another requirement is the size of the photocathode for a single channel, which must match the readout fibers of one scintillator tile. The Hamamatsu R4760 16-channel photomultiplier has been extensively tested for our application (see also \cite{R476ref}). This is a 4~x~4 multichannel PMT with a front face of 70~mm diameter. Each of the 16 channels has a 10 stage dynode chain, but they all share the same voltage divider. The diameter of the photocathode for each channel is 8~mm. Our PMT's fulfill the following requirements: \begin{itemize} \item cathode sensitivity $>$ 45 $\mu$A/lm \item minimum gain at 1000 V: 1~x~10$^6$ \item gain spread between channels, within one PMT assembly, less than a factor of 3 \end{itemize} The crosstalk has been measured to be less than 3\%. \subsection{HV supply and linearity measurement} A high voltage system based on the use of a Cockcroft--Walton generator has been developed \cite{henk_hv}. Power dissipation is negligible compared to that of a resistive voltage divider. The system can be safely operated because of the low voltage input and provides a protection against high currents (light leaks); the maximum anode current is 100$\mu$A. The HV units consist of a microprocessor board and the voltage multiplier boards. The microprocessor performs the HV setting and monitoring. The R4760 PMT operates in the range 800-1200V. Tests showed that the optimum linearity is obtained (at the cost of a slightly lower gain) if the dynode voltage differences are distributed in the proportions 2:1:1:1:1:1:2:2:4:3, starting at the cathode. Since a pulse height measurement is required, good knowledge of the relation between input charge and output signal is necessary. For the linearity measurement we used an LED and a linear neutral density filter. As an example we show in figure~\ref{pmlinea} the deviation from linear behaviour versus anode charge for a HV setting of 1100V. The dotted line shows the linear fit through the first four points, the dashed line represents a polynomial fit through the last four points. \begin{figure}[tbp] \centerline{\psfig{figure=DESY-96-139_6.ps,bbllx=0pt,bblly=150pt,bburx=550pt,bbury=670pt,height=8cm}} \caption{\it Results on the PMT nonlinearity. The line shows the linear fit through the first four points. The dashed line represents a polynomial fit through the last four points.} \label{pmlinea} \end{figure} At an anode charge of 25 pC the measured values are about five percent lower than expected for linear behaviour. These values vary strongly from channel to channel and from PMT to PMT. \section {Readout system} The readout system is a copy of the existing ZEUS calorimeter readout system with some minor modifications \cite{caldwell}. The PMT pulses are amplified and shaped by a pulse shaper circuit mounted at the detector. The shaped pulse is sampled every 96 ns (the bunch crossing rate of the HERA storage ring) and stored in a switched capacitor analog pipeline. After receipt of a trigger from the ZEUS detector, eight samples are transferred from the pipeline to an analog buffer and multiplexed to ADCs. The data are sent to a location outside the detector where the digitisation and signal processing takes place. The modifications consist of upgraded versions of the shaping/amplifier~\cite{iris} and the digital signal processor and a different mechanical layout of the analog front end cards. \section {Results from cosmic ray measurements} A cosmic ray test was performed to measure the light yield of the 576 tiles ( 264 FCAL and 312 RCAL tiles) assembled in 76 cassettes. The trigger system consists of eight cosmic ray telescopes. Each of these consists of two scintillator pads, 20~cm apart, with an area of 12~x~12~cm$^{2}$ which give, together with a third scintillator counter of 240~x~12~cm$^{2}$, a trigger system with a three-fold coincidence. The efficiency of each single telescope was better than 98\%. The readout PMT was a 16-channel R4760 as used in the final presampler design. The setup allows the measurement of 16 channels simultaneously ( e.g. two cassettes with eight tiles each). The trigger rate of the complete setup was about 50 counts/min with about 6 counts/min for a single telescope. An example of the cosmic-test measurement is given in figure~\ref{cos.ps}. To compare the light yield of different tiles, all 16 channels of the PMT were calibrated with a reference tile to correct for differences in quantum efficiency (QE) and gain between the 16 PMT channels. Figure~\ref{fig_tst} shows the mean value for all 576 tiles normalized to one of them. From the RMS value of the distribution we conclude that the responses of all tiles are equal to within 12\%. \begin{figure}[htbp] \centerline{\psfig{figure=DESY-96-139_7.ps,bbllx=45pt,bblly=160pt,bburx=530pt,bbury=650pt,height=5.7cm}} \caption{\it Average pulse height for cosmic muons normalized to one after having corrected for the gain differences between individual PMT channels.} \label{fig_tst} \end{figure} Figure~\ref{fig_npe} shows the mean number of photoelectrons for each tile assuming a QE of 8.5\% which is the minimum value accepted for the presampler PMT's (the mean QE for all channels is 11.6\%). \begin{figure}[htbp] \centerline{\psfig{figure=DESY-96-139_8.ps,bbllx=45pt,bblly=150pt,bburx=530pt,bbury=650pt,height=5.7cm}} \caption{\it Average number of photoelectrons per tile per MIP} \label{fig_npe} \end{figure} \vskip 3cm \section {Calibration tools} \subsection{Minimum--ionising particles in situ} During the operation of ZEUS, halo muons and charged hadrons are used to determine the response to single particles for each individual channel. The high voltage setting common to the sixteen pixels of one R4760 PMT is chosen such that the pixel with the least gain has an average response to minimum--ionising particles which is a factor of ten greater than the RMS noise level of the analog signal--processing front--end electronics (0.05 pC). The pixel--to--pixel gain variation of about a factor of three within one PMT results in a similar variation in the saturation levels. The in situ calibration for the 1995 running period achieved a precision of better than 5\% per tile. \subsection{The radioactive source system} We use an LED/laser system to monitor the gains of the PMT's and a source system to monitor the combined response of tile, fiber and PMT. The response to a $^{60}$Co source provides a relative calibration and quality control of the individual channels of the presampler. The source scans take place during shutdowns of HERA and provide information on the long term behaviour of the light output of the combination of scintillator and wavelength-shifting fiber. \begin{figure}[htbp] \centerline{\psfig{figure=DESY-96-139_9.eps,bbllx=203pt,bblly=0pt,bburx=0pt,bbury=608pt,angle=270,width=15cm}} \caption{\it Example of a cassette with 9 scintillating tiles and source tube glued on top, connected to source scanning system.} \label{source_layout} \end{figure} Brass tubes with 2.5 mm outer diameter and 0.2 mm wall thickness run over the full length of the cassette, positioned in the middle (figure~\ref{source_layout}). They guide the pointlike (0.8~mm diameter, 1~mm length, 74 MBq) source. The source is driven in 2~mm steps via a 1.2~mm diameter steel wire by a stepper motor~\cite{source_ref} controlled by a PC. The PMT currents are integrated with a time constant of 24 ms and read into a 16-channel 12-bit ADC card. Figure~\ref{x3} shows as an example the superposition of the responses of the tiles within one cassette as a function of the location of the source. The different heights of the maxima are mainly due to the different gains of the 16 channels of the R4760 PMT, all supplied with the same high voltage. The $^{60}$Co source is not collimated, as can be seen in the shape of the individual peaks. \begin{figure}[htbp] \centerline{\psfig{figure=DESY-96-139_10.eps,width=8cm,height=10cm}} \caption{\it Responses of the scintillating tiles within one cassette to a $^{60}$Co source. The step width is approximately 2 mm.} \label{x3} \end{figure} \clearpage \section {Beam test results} The influence of material in front of the ZEUS calorimeter on its energy measurement has been studied previously in several test beam runs with the ZEUS forward calorimeter (FCAL) prototype \cite{prot}. The corrections to the calorimetric measurements that can be derived from presampling measurements have been studied in subsequent test periods for both hadrons and electrons~\cite{marcel}. In the following we summarize the most recent results obtained for electrons with the final presampler design~\cite{adi}. \begin{figure}[tbp] \centerline{\psfig{figure=DESY-96-139_11.eps,bbllx=30pt,bblly=440pt,bburx=565pt,bbury=700pt,height=6cm}} \caption{\it Experimental setup of the FCAL prototype and presampler in CERN test beam. The presampler is mounted directly on the FCAL frontplate.} \label{preprot} \end{figure} \subsection{Overview} \normalsize The presampler prototype consists of an array of 4 x 4 scintillator tiles covering an area of 80 x 80 cm$^2$. As shown in figure~\ref{preprot} it is positioned directly in front of the ZEUS FCAL prototype which has the same lateral size. The depth of the calorimeter is 7 interaction lengths \cite{prot}. Beam tests were performed in the X5 test beam of the CERN SPS West Area. The prototype presampler detector is read out via an R4760 multichannel photomultiplier using the Cockcroft--Walton HV system. Furthermore, the final readout electronics was used for both the presampler and the FCAL prototype modules. The uranium radioactivity was used to set the relative gains of the calorimeter phototubes and 15 GeV electrons served to set the energy scale. Muons were used to calibrate the presampler. The combined response of the presampler and calorimeter was determined for electrons in the energy range from 3-50 GeV. The amount of material installed in front of the presampler varied between 0 and 4 radiation lengths (X$_0$) of aluminium. During these studies, the position of both calorimeter and presampler relative to the beam was fixed. A delay wire chamber allowed the determination of the impact point of the beam particles with an accuracy of 0.5~mm. Most of the data were recorded with a defocussed beam about 10 cm in diameter, facilitating studies of uniformity and position dependence of the energy correction algorithms. \subsection{ The uniformity of the presampler response to muons } Figure~\ref{f9-1} shows the mean presampler response to 75 GeV muons. The position information was provided by the delay wire chamber. The presampler signals for the incident muons are normalized to the response at the center of the tile and averaged over uniformly populated rectangles of 90~x~5~mm$^2$. The nonuniformity in the sum of the two bordering tiles is a few percent in the regions of the fibers. In the horizontal direction (figure~\ref{f9-1}b) the tiles are mounted within one cassette with no gaps between them. In the vertical coordinate (figure~\ref{f9-1}c) a signal drop is observed between the cassettes due to the 1.4 mm gap between the scintillator tiles. The nonuniformity averaged over the surface of a tile is less than 1\%. \begin{figure}[hb] \begin{picture}(400,170)(0,0) \put(0,18) {\psfig{figure=DESY-96-139_12.ps,bbllx=6pt,bblly=40pt,bburx=580pt,bbury=630pt,height=4.7cm}} \put(130,0) {\psfig{figure=DESY-96-139_13.eps,height=6cm}} \put(285,0){\psfig{figure=DESY-96-139_14.eps,height=6cm}} \put(10,134){\large{a)}} \put(160,80){\large{b)}} \put(320,80){\large{c)}} \end{picture} \caption{\it The muon response uniformity of the presampler near tile borders.\hspace{2cm} a) A sketch of scanning region, b) a horizontal scan (* represents the response of tile 6, $\circ$ that of tile 7 and $\bullet$ the sum of both), c) a vertical scan, perpendicular to the embedded fibers. The dashed lines indicates the fiber positions. The data in both plots are averaged over a uniformly populated rectangle 90~x~5~mm$^2$ in area with the long side perpendicular to the scanning direction.} \label{f9-1} \end{figure} \subsection{The presampler response to electrons} The electron beam used for the energy correction studies was 1 cm wide and 10 cm high, centered horizontally within one FCAL module and vertically on one of the 20~x~5~cm$^{2}$ electromagnetic sections. The presampler signal ($E_{pres}$) was obtained by summing all 16 tiles in order to be sure to get the entire signal and because the electronic noise contribution was negligible. The signal from each tile was normalized to its average response to muons, resulting in units we refer to as ``MIP''. Aluminium plates of 3 cm thickness were used as the absorber material. In the following three such plates together are referred to as one radiation length, an approximation which is accurate to 1\%. As examples of the calorimeter and presampler signal spectra we show in figure~\ref{calvpres} the energy distributions measured with the calorimeter ($E_{cal}$) and the signal in the presampler for 25 GeV electrons for aluminium absorber thicknesses ranging between 0 and 4 X$_0$. The mean value for $E_{cal}$ decreases by more than 20\% but the shapes of the distributions remain approximately gaussian. The resolution deteriorates substantially in the presence of more than 2 X$_0$ of absorber material. \begin{figure}[tbp] \centerline{\psfig{figure=DESY-96-139_15.ps,bbllx=10pt,bblly=130pt,bburx=550pt,bbury=700pt,height=12cm}} \caption{\it Signal distributions in the calorimeter and presampler for 25 GeV electrons having passed through 0, 1, 2, 3, and 4 radiation lengths of aluminium absorber.} \label{calvpres} \end{figure} \clearpage Table 1 shows the relative calorimeter signal loss and the average and RMS values of the presampler signal spectra for the full range of electron energies and absorber thicknesses. The uncertainties presented are dominated by the statistical precision.\\*[3mm] \begin{tabular}{|c|c|rcl|rcl|rcl|} \hline Energy & Absorber & \multicolumn{3}{|c|}{Rel. Energy Loss} & \multicolumn{3}{|c|}{Presampler Avg} & \multicolumn{3}{|c|}{Presampler RMS} \\ (GeV) & ($X_0$) & \multicolumn{3}{|c|}{(\%)} & \multicolumn{3}{|c|}{(MIP)} & \multicolumn{3}{|c|}{(MIP)} \\ \hline 3& 1& \hspace*{1mm} 7.8& $\pm$ &1.0& 5.5& $\pm$ &0.1& \hspace*{2mm} 3.7& $\pm$ &0.1\\ \cline{3-11} & 2& \hspace*{1mm} 18.0& $\pm$ &1.0& 10.3& $\pm$ &0.2& \hspace*{2mm} 5.3& $\pm$ &0.1\\ \cline{3-11} & 3& \hspace*{1mm} 32.8& $\pm$ &1.0& 12.0& $\pm$ &0.2& \hspace*{2mm} 5.6& $\pm$ &0.1\\ \cline{3-11} & 4& \hspace*{1mm} 49.4& $\pm$ &1.0& 12.1& $\pm$ &0.2& \hspace*{2mm} 5.5& $\pm$ &0.1\\ \hline 5& 1& \hspace*{1mm} 6.4& $\pm$ &0.6& 6.2& $\pm$ &0.2& \hspace*{2mm} 4.1& $\pm$ &0.1\\ \cline{3-11} & 2& \hspace*{1mm} 13.0& $\pm$ &0.7& 12.8& $\pm$ &0.2& \hspace*{2mm} 6.5& $\pm$ &0.2\\ \cline{3-11} & 3& \hspace*{1mm} 27.4& $\pm$ &0.7& 16.9& $\pm$ &0.3& \hspace*{2mm} 7.3& $\pm$ &0.2\\ \cline{3-11} & 4& \hspace*{1mm} 42.5& $\pm$ &0.7& 19.5& $\pm$ &0.3& \hspace*{2mm} 7.1& $\pm$ &0.2\\ \hline 10& 1& \hspace*{1mm} 4.8& $\pm$ &0.5& 7.7& $\pm$ &0.2& \hspace*{2mm} 5.0& $\pm$ &0.1\\ \cline{3-11} & 2& \hspace*{1mm} 10.1& $\pm$ &0.5& 18.2& $\pm$ &0.4& \hspace*{2mm} 8.9& $\pm$ &0.3\\ \cline{3-11} & 3& \hspace*{1mm} 20.3& $\pm$ &0.5& 27.4& $\pm$ &0.4& \hspace*{2mm} 10.9& $\pm$ &0.3\\ \cline{3-11} & 4& \hspace*{1mm} 34.6& $\pm$ &0.7& 32.7& $\pm$ &0.6& \hspace*{2mm} 11.3& $\pm$ &0.4\\ \hline 15& 1& \hspace*{1mm} 3.5& $\pm$ &0.4& 7.9& $\pm$ &0.2& \hspace*{2mm} 5.2& $\pm$ &0.2\\ \cline{3-11} & 2& \hspace*{1mm} 7.8& $\pm$ &0.4& 20.4& $\pm$ &0.5& \hspace*{2mm} 9.6& $\pm$ &0.4\\ \cline{3-11} & 3& \hspace*{1mm} 17.3& $\pm$ &0.4& 34.9& $\pm$ &0.6& \hspace*{2mm} 13.8& $\pm$ &0.4\\ \cline{3-11} & 4& \hspace*{1mm} 29.6& $\pm$ &0.6& 46.0& $\pm$ &0.8& \hspace*{2mm} 14.6& $\pm$ &0.6\\ \hline 25& 1& \hspace*{1mm} 2.7& $\pm$ &0.3& 9.8& $\pm$ &0.2& \hspace*{2mm} 6.5& $\pm$ &0.2\\ \cline{3-11} & 2& \hspace*{1mm} 5.9& $\pm$ &0.3& 26.6& $\pm$ &0.4& \hspace*{2mm} 12.8& $\pm$ &0.3\\ \cline{3-11} & 3& \hspace*{1mm} 15.4& $\pm$ &0.3& 47.3& $\pm$ &0.5& \hspace*{2mm} 17.9& $\pm$ &0.4\\ \cline{3-11} & 4& \hspace*{1mm} 24.4& $\pm$ &0.3& 66.1& $\pm$ &0.6& \hspace*{2mm} 20.9& $\pm$ &0.4\\ \hline 50& 1& \hspace*{1mm} 1.6& $\pm$ &0.2& 12.0& $\pm$ &0.3& \hspace*{2mm} 7.9& $\pm$ &0.2\\ \cline{3-11} & 2& \hspace*{1mm} 4.6& $\pm$ &0.2& 35.6& $\pm$ &0.6& \hspace*{2mm} 16.4& $\pm$ &0.4\\ \cline{3-11} & 3& \hspace*{1mm} 11.5& $\pm$ &0.5& 72.7& $\pm$ &2.1& \hspace*{2mm} 27.7& $\pm$ &1.5\\ \cline{3-11} & 4& \hspace*{1mm} 19.8& $\pm$ &0.3&108.1& $\pm$ &1.2& \hspace*{2mm} 31.6& $\pm$ &0.8\\ \hline \end{tabular} \vskip 5mm Table 1: {\it The relative decrease in the calorimeter signal and the average and RMS values of the presampler signal spectra for each electron energy and each aluminium absorber thickness used in the test beam studies. The uncertainties shown are dominated by the statistical precision.} \clearpage \subsubsection {Contribution of backscattering to the presampler signal} The comparison of presampler signal spectra from incident muons with those of incident electrons allowed us to estimate the contribution of backscattering from the electromagnetic showers in the uranium calorimeter. Figure~\ref{backscat} shows this comparison for 5, 15, 25 and 50 GeV electrons. We determine the average relative increase to be 1.45, 1.65, 1.80, and 2.16 respectively, to an accuracy of 2\%. By adding absorber upstream of the presampler and displacing both absorber and presampler several meters upstream to measure the decreased contribution from backscattering, we ascertained that the backscattering contribution is not increased by the presence of the absorber in front of the presampler. Thus we can be sure that the backscattering contribution remains at the level of 1 MIP and can be neglected at the level of 10\% of the size of the signals we use for the electromagnetic energy correction. We also measured the backscattering contribution from hadronic showers and found for 15 and 75 GeV incident pions values for the average relative increase less than 1.5 and 2.0 respectively. \begin{figure}[tbp] \centerline{\psfig{figure=DESY-96-139_16.ps,bbllx=50pt,bblly=270pt,bburx=530pt,bbury=540pt,height=8cm}} \caption{\it A comparison of the presampler signal spectra for 5, 15, 25, and 50 GeV electrons to that for muons. The smooth curve shows the response to muons. The presampler signal has been normalized to the average value of the muon spectrum. Each spectrum contains 2000 entries.} \label{backscat} \end{figure} \subsubsection{Electron energy correction} Figure~\ref{f9-4} shows the correlation between the presampler signal (normalized to the average signal of a minimum--ionising particle) and the calorimeter signal for 25 GeV electrons as a function of the amount of absorber material. \begin{figure}[tbp] \centerline{\psfig{figure=DESY-96-139_17.ps,bbllx=25pt,bblly=270pt,bburx=535pt,bbury=545pt,height=8cm}} \caption{\it Calorimeter versus presampler response for 25 GeV electrons and absorber material ranging from 1 to 4 X$_0$. The line represents the fit to the data according formula (1).} \label{f9-4} \end{figure} We have considered a variety of parametrisations for the relationship between calorimeter and presampler responses. In the well--defined environment of a test beam the correction is straightforward and depends on the incident energy and the amount of absorber material, both of which are precisely known. \begin{figure}[tbp] \vskip -2cm \centerline{\psfig{figure=DESY-96-139_18.ps,bbllx=50pt,bblly=265pt,bburx=530pt,bbury=540pt,height=8cm}} \caption{\it Average calorimeter response normalized to the electron energy versus the electron energy before and after correction} \label{f9-2} \end{figure} In a detector environment the amount of absorber material in front of the calorimeter is not uniformly distributed, arising from cables, support structures, etc. One can, however, identify regions where the average amount of absorber material is roughly known. For this reason we show here the result obtained with one set of correction constants common to the 1 and 2 X$_0$ data set and one for the 3 and 4 X$_0$ data set. The relation between the measured mean values of $E_{cal}$ and $E_{pres}$ has been parametrised in a linear approximation: \begin{equation} \label{eq:etrue} E_{cal}= a_0 + a_1 E_{pres} \end{equation} The result for the two data sets for 25 GeV electrons is shown in figure~\ref{f9-4}. The parameters $a_i$ depend on the amount of material and on the electron beam energy. This correction algorithm allows for a linear energy dependence of the parameters $a_i$: $a_i = \alpha_i +\beta_i E_{beam}$. We neglect the dependence on the amount of absorber material in order to estimate the success of the algorithm when the amount of absorber varies within the data sample. The parameters $\alpha_i$ and $\beta_i$ are determined by minimising the difference of the beam energy and the corrected calorimeter signal, The results for the corrected calorimeter response for electrons in the energy range 3-50 GeV, are shown in figure~\ref{f9-2}. This procedure provides a correction accurate to 3\% for the energy range studied here, but for an overcorrection of about 10\% for the 3~X$_0$ data points at low energy. For electron energies greater than 5 GeV and for absorber thickness less than 2 X$_0$, the values relevant to the operation of the ZEUS detector, this simple correction algorithm yields a systematic precision of 2\%. The improvement in the energy resolution as well as in the energy scale is shown in figure~\ref{mixture}. The energy distribution of 25 GeV electrons for a merged 1-4 X$_0$ data set is shown before and after correction. \begin{figure}[htbp] \centerline{\psfig{figure=DESY-96-139_19.ps,bbllx=25pt,bblly=270pt,bburx=530pt,bbury=540pt,height=6.5cm}} \caption{\it Reconstructed energy distributions for 25 GeV electrons for a mixture of the 1-4 X$_0$ data before and after application of the correction algorithm} \label{mixture} \end{figure} \section {Summary and conclusions} We have designed, built, installed and operated a scintillator-tile presampler for the forward and rear calorimeter of the ZEUS detector. The scintillation light from the tiles is collected by wavelength-shifting fibers and guided to multi-anode photomultipliers via clear fibers. The signals are shaped, sampled and pipelined in the manner employed for the calorimeter itself, easing the integration of the presampler in the ZEUS data acquisition system. The performance of the tiles and fiber readout were monitored with a cosmic--ray telescope and with collimated sources during production. The single-particle detection efficiency is greater than 99\% and the response uniformity over the area of each of the 576 20~x~20~cm$^{2}$ tiles is better than 5\%. An LED flasher system and scans with radioactive sources have proven useful diagnostic tools since the installation of the presampler. The in situ calibration with minimum--ionising particles during the 1995 data--taking period achieved a precision better than 5\% per tile. Test beam studies of a presampler prototype with the final geometry and readout in combination with a prototype of the ZEUS forward calorimeter verified the efficiency and uniformity results and allowed the determination of backscattering contributions. Tests with electrons were performed in the energy range 3--50 GeV with 0--4 radiation lengths of aluminium absorber. These studies have proven the feasibility of an electron energy correction accurate to better than 2\% in the energy range and for the configuration of inactive material relevant to the ZEUS detector. \section {Acknowledgements} We would like to thank the technical support from the institutes which collaborated on the construction of the presampler, in particular W. Hain, J. Hauschildt, K. Loeff\-ler, A. Maniatis, H.--J. Schirrmacher (DESY), W. Bienge, P. Pohl, K.--H. Sulanke (DESY--IfH Zeuthen), M. Gospic, H. Groenstege, H. de Groot, J. Homma, I. Weverling, P. Rewiersma (NIKHEF), R. Mohrmann, H. Pause, W. Grell (Hamburg), M. Riera (Madrid), R. Granitzny, H.--J. Liers (Bonn). We are also grateful for the hospitality and support of CERN, in particular the help of L. Gatignon is much appreciated. Finally we would like to thank R. Klanner for his enthusiastic support during all phases of the project.
proofpile-arXiv_065-453
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Brief introduction} In the present talk we discuss the spin-dependent structure functions (SF) of nuclei and their relation to those of nucleon. Our main focus will be the deuteron, which we study in detail in the covariant Bethe-Salpeter formalism. Why it so important and interesting to study the nuclear effects in the SFs? First, nuclei are the only source of the experimental information about neutron SFs, including the spin-dependent ones. To obtain this information, it is important to understand how nucleons are bound in the nuclei and how this binding affects their SFs. An accurate method to extract the neutron SFs from the nuclear data must be an essential part of the consistent analysis of the nucleon SFs. Second, a physics governing the processes with the participation of the nuclei is extremely interesting by itself. For instance, a spin-1 nucleus, such as a deuteron, has extra spin-dependent SFs than nucleons, i.e. $b_{1,2}^D$. Another example, nuclei as a slightly relativistic and weakly bound systems allows for more progress than the hadrons, in studying the covariant bound state problem. In certain situations the covariant approach gives results noticeably different from the nonrelativistic ones. For the spin-dependent SFs such situation is a calculation of the $b_{1,2}^D$. And third, our interest in the study of the reactions with the deuteron is in part motivated by the future and ongoing experiments. In particular, very recently we started a study of the chiral-odd SF $h_1^D$. \section{Spin-dependent SF of nucleon, $g_1^N$. } For recent reviews about the nucleon spin-dependent SFs see refs.~\cite{jaffer,rev2}. \subsection{Basic formulae} The differential cross section for the polarized electron-nucleon scattering has the form: \begin{eqnarray} \frac{d^2\sigma}{d\Omega dE'} = \frac{\alpha^2E'}{2mq^4E} L^{\mu\nu}W_{\mu\nu}, \label{crosec} \end{eqnarray} where $\alpha = e^2/(4\pi)$, $q=(\nu,0,0,-\sqrt{\nu^2+Q^2})$ is the momentum transfer, $ Q^2=-q^2$, $m$ is the nucleon mass, $E(E')$ is the energy of the incoming (outgoing) electron, $L^{\mu\nu}$ and $W_{\mu\nu}$ are the leptonic and hadronic tensors. The most general expression of $W_{\mu\nu}$ is \begin{eqnarray} &&\!\!\!\!\!\!\!\!\!\!\!\!\!\!\! W_{\mu\nu}^N(q,p) =\label{htenn} \\ &&\!\!\!\!\!\!\!\!\! \left ( -g_{\mu\nu} +\frac{q_\mu q_\nu}{q^2}\right ) F_1^N(x_N,Q^2) + \left ( p_{\mu} - q_\mu \frac{pq}{q^2} \right ) \left ( p_{\nu} - q_\nu \frac{pq}{q^2} \right ) \frac{F_2^N(x_N,Q^2)}{pq} \nonumber\\ +&&\!\!\!\!\!\!\!\!\! \frac{im}{pq} \epsilon_{\mu\nu\alpha\beta} q^\alpha \left \{ S^\beta \left (g_1^N(x_N,Q^2) + g_2^N(x_N,Q^2) \right ) -p^\beta \frac{(Sq)}{pq}g_2^N(x_N,Q^2) \right \} , \nonumber \end{eqnarray} where $x_N = Q^2/(2pq)$ (in the rest frame of the nucleon $x_N = Q^2/(2m\nu)$) and $S$ is the nucleon spin. In accordance to the ideology of the quark-parton model, the SFs $F_{1,2}$ and $g_1$ are proportional to the appropriate quark distributions on a part of the total longitudinal momentum of the nucleon, $x$. For instance, if we denote the net spin carried by quarks as $\Delta q(x,Q^2)$ ($q = u,d,s$) and introduce the following combinations: \begin{eqnarray} &&\Delta q_3(x,Q^2) \equiv \Delta u(x,Q^2) - \Delta d(x,Q^2) , \label{su33}\\ &&\Delta q_8(x,Q^2) \equiv \Delta u(x,Q^2) + \Delta d (x,Q^2) -2\Delta s(x,Q^2) , \label{su38}\\ &&\Delta \Sigma (x,Q^2) \equiv \Delta u(x,Q^2) + \Delta d(x,Q^2)+ \Delta s(x,Q^2), \label{su30} \end{eqnarray} then the SFs of the proton(neutron) can be written as: \begin{eqnarray} g_1^{p,n}(x,Q^2) = \pm \frac {1}{12}\Delta q_3(x,Q^2) + \frac{1}{36} \Delta q_8(x,Q^2) +\frac{1}{36}\Delta \Sigma(x,Q^2) . \label{g1pm} \end{eqnarray} Important objects of the study of quark structure of the hadrons are the so-called sum rules for the SFs. The sum rules relate the moments of the SFs to the fundamental (or sometimes not very fundamental) constants of the theory. Integrating eq.~(\ref{g1pm}), the first moments of the proton (neutron) structure functions can be written in self-explaining notation: \begin{eqnarray} S^{p(n)} &\equiv& \int\limits_0^1 dx g_1^{p(n)}(x,Q^2) =\label{ig1p}\\ &&\!\!\!\! \frac {1}{12} \left( 1 -\frac{\alpha_s}{\pi}+ {\small \ldots} \right) \left (\pm \Delta q_3 + \frac{1}{3} \Delta q_8 \right ) + \frac{1}{9} \left( 1 -\frac{\alpha_s}{3\pi}+ {\small \ldots} \right)\Delta \Sigma , \label{ignuc}\end{eqnarray} where the perturbative QCD corrections, to order ${\cal O}(\alpha_s)$, are also presented. From the current algebra for asymptotic integrals we have ($Q^2 \to \infty $): \begin{eqnarray} \Delta q_3 = 1.257\pm 0.003 , \quad \Delta q_8 = 0.59 \pm 0.02\; (?). \label{isu38} \end{eqnarray} The first constant is from the weak decay of the neutron and the second constant is from the decay of the hyperon. The ``?'' mark is due to the residual questions about SU(3). The Bjorken sum rule is the most fundamental relation: \begin{eqnarray} S^p - S^n &=& \frac {1}{6} \left( 1 -\frac{\alpha_s}{\pi}+ \cdots \right) \Delta q_3, \label{bsr}\end{eqnarray} which numerically gives $0.187 \pm 0.003$ at $Q^2 = 10$~GeV$^2$ and $0.171 \pm 0.008$ at $Q^2 = 3$~GeV$^2$. The Ellis-Jaffe sum rule is not so fundamental. Assuming that $\Delta s = 0$ and, therefore, $\Delta \Sigma = \Delta q_8 \simeq 0.6$, we get at $Q^2=10$~GeV$^2$ ($3$~GeV$^2$): \begin{eqnarray} S^p_{EJ} &=& \phantom{-}0.171\pm 0.004 \quad (\phantom{-}0.161 \pm 0.004) ,\label{ejp}\\ S^n_{EJ} &=& -0.014\pm 0.004 \quad (-0.010 \pm 0.004).\label{ejn} \end{eqnarray} The spin-dependent SFs, $g_1$, allow also to study spin content of the hadrons. Indeed, using eqs.~(\ref{isu38}) and experimental values of $S^{p,n}$ (a fraction of) the nucleon spin carried by quarks, $\Delta \Sigma$, can be determined. The total angular momentum (spin) of the nucleon consists not only of $\Delta\Sigma$, but also: \begin{eqnarray} \frac{1}{2} = \frac{1}{2}\Delta \Sigma + \Delta G + L_z^g +L_z^G, \label{spincont} \end{eqnarray} where $\Delta G$ is the gluon spin contribution, $L_z^{q(G)}$ is the quark (gluon) orbital angular momentum contribution. In naive quark model $\Delta \Sigma =1$ and others are zeros. In the relativistic quark model $\Delta \Sigma =0.75$ and $L_z^q=0.125$ and others are zeros. From the current algebra $\Delta \Sigma \approx 0.6 \pm 0.1$, others are unknown. \subsection{Experiments for $g_1^N$} Both the SFs and the sum rules are the subject of intensive experimental studies in recent years. Table~I presents measurements by various experimental groups. \begin{center} {\small {\normalsize \bf Table I.} \vskip .25cm \begin{tabular}{|c|c|c|c|c|} \hline Experiment & Year & Target & $\sim Q^2$~GeV$^2$ & $S^{Target} $ \\ \hline \hline E80/E130 & 1976/1983 & p & 5 & 0.17 $\pm$ 0.05 \\ EMC & 1987 & p & 11 & 0.123 $\pm $ 0.013 $\pm$ 0.019\\ SMC & 1993 & d & 5 & 0.023 $\pm $ 0.020 $\pm$ 0.015\\ SMC & 1994 & p & 10 & 0.136 $\pm $ 0.011 $\pm$ 0.011\\ SMC & 1995 & d & 10 & 0.034 $\pm $ 0.009 $\pm$ 0.006\\ E142 & 1993& n ($^3He$)& 2 & -0.022 $\pm $ 0.011\\ E143 & 1994& p & 3 & 0.127 $\pm $ 0.004 $\pm$ 0.010\\ E143 & 1995& d & 3 & 0.042 $\pm $ 0.003 $\pm$ 0.004\\ HERMES& 1996& n ($^3He$)& 3 & -0.032 $\pm $ 0.013 $\pm$ 0.017\\ \hline \end{tabular} } \end{center} \vskip .2cm From the SMC and E143 data the Bjorken sum rule is: \begin{eqnarray} \left (S^p - S^n \right )_{SMC} &\approx& 0.199 \pm 0.038 \quad{\rm at} \quad Q^2 = 10\quad{\rm GeV}^2,\label{bsre10}\\ \left (S^p - S^n \right) _{E143} &\approx& 0.163 \pm 0.010 \pm 0.016 \quad{\rm at} \quad Q^2 = 3\quad{\rm GeV}^2,\label{bsre3} \end{eqnarray} i.e. the sum rule is confirmed with 10 \% accuracy. From Table~I it is clear that the Ellis-Jaffe sum rules are broken. As to the spin content, (\ref{spincont}), only one piece, $\Delta \Sigma$, can be extracted from the data for the integrals of SFs. The world data from Table~I gives: \begin{eqnarray} \Delta \Sigma \approx 0.3\pm 0.1, \label{spinconte} \end{eqnarray} which is larger than the first result of EMC, $\Delta \Sigma = 0.12\pm 0.094\pm 0.138\approx 0$, but still lower than quark model estimates. In addition to the perturbative corrections in eq.~(\ref{ig1p}), various other corrections, such as the kinematic mass corrections, $\sim m^2/Q^2$ and higher twist corrections, $\sim 1/Q^2$, are discussed. \section{Nucleons and nuclei} Note that actual data for the neutron is not presented in Table~I, only the data for lightest nuclei. A simple formula is used to obtain $g_1^n$ from the combined proton and deuteron data: \begin{eqnarray} g_1^D = \left (1-\frac{3}{2}w_D\right )\left (g_1^p+g_1^n\right), \label{simple} \end{eqnarray} where $w_D$ is the probability of the $D$-wave state in the deuteron. Depending on the model, $w_D = 0.04 - 0.06$. Similarly, the neutron SF is obtained from the $^3He$ data: \begin{eqnarray} g_1^{^3He} = \left ( P_S + \frac{1}{3}P_{S'} -P_D \right ) g_1^n +\left (\frac{2}{3}P_{S'} - \frac{2}{3}P_D \right ) g_1^p, \label{simple1} \end{eqnarray} where $P_S$, $P_{S'}$ and $P_D$ are weights of the $S$, $S'$ and $D$ waves in $^3He$, respectively. Typical (model-dependent) values of these weights are $P_S\approx 0.897$, $P_{S'}=0.017$ and $P_D=0.086$. It is important to realize that the real connection between the nucleon and nuclear SFs is more complex than given by formulae like eq.~(\ref{simple}) and (\ref{simple1}). Studies of the last decade show importance of the proper separation of the binding, Fermi motion and the off-mass-shell effects in the procedure of extracting the neutron SFs from the nuclear data (see refs.~\cite{amb,unfo,mtamb} and references therein). However, effects of the Fermi motion are sometimes estimated by the experimental groups, other effects are always neglected. Such a way of action can be phenomenologically more or less safe at the present level of accuracy of the experiments, but not in general. The deuteron is the most appropriate target to study the neutron SFs, since it has a well-known structure and well-studied wave function or relativistic amplitude. Besides all other effects such as meson exchanges, binding of the nucleons, off-mass-shell corrections, shadowing, etc, are minimal. Even in the case of $^3$He the situation is known to be different~\cite{kubj,fshe}. Indeed, eq.~(\ref{simple1}) or even more sophisticated convolution formula violate the fundamental Bjorken sum rule for the $^3$He-$^3$H pair at the 3-5\% level, which is a serious indication of other degrees of freedom involved in the process. Once again, this fact is completely ignored by the experimental groups reporting the results for the neutron SFs from experiments with $^3$He. In what follows we consider the nuclear effects in the spin-dependent SFs of the deuteron. Results of our studies make us certain that an accurate extraction of the {\em neutron } spin structure function, $g_1^n$, is possible. Considering nuclei as a complex system of interacting nucleons and mesons, we calculate the nuclear SFs in terms of the structure functions of its constituents, nucleons and mesons, and in the Bethe-Salpeter formalism for the deuteron amplitude. For the spin-independent SFs, $F^D_2$, the mesonic contributions to the SF is important (although quite small) for the consistency of the approach, since the mesons carry a part of the total momentum of the nuclei (see~\cite{uk} and references therein). However, for the spin-dependent SFs explicit contribution of mesons is not important. Rather, their presence manifests via binding of nucleons in nuclei. This is why we consider only nucleon contributions to the spin-dependent SFs. We start with the general form of the hadron tensor of the deuteron with the total angular momentum projection, $M$, keeping only leading twist SFs: \begin{eqnarray} W_{\mu\nu}^D(q,P_D,M) &=& \left ( -g_{\mu\nu} +\frac{q_\mu q_\nu}{q^2}\right ) F_1^D(x_D,Q^2,M) + \label{htend} \\ && \left ( P_{D\mu} - q_\mu \frac{P_Dq}{q^2} \right ) \left ( P_{D\nu} - q_\nu \frac{P_Dq}{q^2} \right ) \frac{F_2^D(x_D,Q^2,M)}{P_Dq} \nonumber\\ && +\frac{iM_D}{P_Dq} \epsilon_{\mu\nu\alpha\beta} q^\alpha S_D^\beta(M) g_1^D(x_D,Q^2), \nonumber \end{eqnarray} where $x_D = Q^2/(2P_Dq)$ (in the rest frame of the deuteron $x_D = Q^2/(2M_D\nu)$), $S_D(M)$ is the deuteron spin and $F_{1,2}^D$ and $g_1^D$ are the deuteron SFs. Averaged over $M$ this expression leads to the well-known form of the spin-independent hadron tensor which is valid for hadron with any spin: {\small\begin{eqnarray} && W_{\mu\nu}^D(q,P_D) = \frac{1}{3}\sum_M W_{\mu\nu}^D(q,P_D,M) \label{av}\\ &=&\!\!\!\! \left ( -g_{\mu\nu} +\frac{q_\mu q_\nu}{q^2}\right ) F_1^D(x_D,Q^2) + \left ( P_{D\mu} - q_\mu \frac{P_Dq}{q^2} \right ) \left ( P_{D\nu} - q_\nu \frac{P_Dq}{q^2} \right ) \frac{F_2^D(x_D,Q^2)}{P_Dq}, \nonumber \end{eqnarray} } where $F_{1,2}^D(x_D,Q^2)$ are the result of averaging of the SF $F_{1,2}^D(x_D,Q^2,M)$. To separate $g_1^D$ we can use one of the following projectors~\cite{ukk}: \begin{eqnarray} R^{(1)}_{\mu\nu} \equiv i\epsilon_{\mu\nu\alpha\beta}q^\alpha S^\beta_D(M), \quad R^{(2)}_{\mu\nu} \equiv \frac{i (S_D(M)q)}{P_Dq} \epsilon_{\mu\nu\alpha\beta}q^\alpha P_D^\beta . \label{aw2} \end{eqnarray} In the limit $Q^2/\nu^2 \to 0$: \begin{eqnarray} g_1^D = \frac{R^{(1)\mu\nu}W_{\mu\nu}^D }{2\nu} = \frac{R^{(2)\mu\nu}W_{\mu\nu}^D }{2\nu} . \label{exg1} \end{eqnarray} The nucleon contribution to the deuteron SFs is presented by the triangle graph, written in terms of the Bethe-Salpeter amplitude of the deuteron~\cite{uk,ukk}: \let\picnaturalsize=N \def3.0in{2.10in} \defh-1d.eps{tre-dia.eps} \ifx\nopictures Y\else{\ifx\epsfloaded Y\else\input epsf \fi \let\epsfloaded=Y \centerline{\ifx\picnaturalsize N\epsfxsize 3.0in\fi \epsfbox{h-1d.eps}}}\fi \noindent where $\hat W$ is the appropriate operator, describing the scattering on the constituent nucleon. Neglecting small correction due to the ``nucleon deformation''~\cite{mst} it can be written down as: \begin{eqnarray} &&\hat W^N_{\mu\nu}(q,p)=\hat W_{\{\mu\nu\}}(q,p) + \hat W_{[\mu\nu]}(q,p) \label{deftens}\\[3mm] &&\hat W_{\{\mu\nu\}}(q,p) = \frac{\hat q}{2pq}W^N_{\mu\nu}(q,p), \label{ops}\\ &&\hat W_{[\mu\nu]}(q,p) = \frac{ i}{2pq} \epsilon_{\mu\nu\alpha\beta} q^{\alpha} \gamma^\beta \gamma_5 g_1^N (q,p) , \label{opa} \end{eqnarray} where $\{\ldots \}$ and $[\ldots ]$ denote symmetrization and antisymmetrization of indices, respectively, $W_{\{\mu\nu\}}^N$ is the spin-independent part of the hadron tensor of the nucleon and $g_1^N(q,p)=g_1^N(x,Q^2)$ is the spin-dependent nucleon SF. The explicit expressions of the deuteron SFs in terms of the Bethe-Salpeter amplitude, $\Psi_M(p_0,{\bf p})$, are given by: \begin{eqnarray} F_2^D(x_N,Q^2,M) &=& i\int \frac{d^4p}{(2\pi)^4} {F_2^N} \left( \frac{x_N m}{p_{10}+p_{13}}, Q^2\right)\label{f2m} \\ &&\frac{ {\sf Tr}\left\{ \bar\Psi_M(p_0,{\bf p})(\gamma_0+\gamma_3) \Psi_M(p_0,{\bf p}) (\hat p_2-m) \right \}}{2M_D}, \nonumber \\[2mm] g_1^D(x_N,Q^2) &=& i \int \frac{d^4p}{(2\pi)^4} {g_1^N}\left( \frac{x_N m}{p_{10}+p_{13}}, Q^2\right) \label{g1m}\\ && \frac{\left. {\sf Tr}\left\{ \bar\Psi_M(p_0,{\bf p})(\gamma_0+\gamma_3)\gamma_5 \Psi_M(p_0,{\bf p}) (\hat p_2-m) \right \}\right |_{M=1}}{2(p_{10}+p_{13})}, \nonumber \end{eqnarray} where $p_{10}$ and $p_{13}$ are the time and 3-rd components of the struck nucleon momentum. Averaging over the projection $M$ has not been done in eq.~(\ref{f2m}), since we use the present form later to calculate the SF $b_{1,2}^D$. Then two independent ``SFs'', with $M = \pm 1$ and $M=0$ are obtained: \begin{eqnarray} &&F_2^D(x_N,Q^2) = \frac{1}{3} \sum_{M=0,\pm 1} F_2^D(x_N,Q^2,M), \label{f2}\\[1mm] && F_2^D(x_N,Q^2,M=+1) = F_2^D(x_N,Q^2,M=-1).\label{pm} \end{eqnarray} A method to calculate numerically expressions like (\ref{f2m}) and (\ref{g1m}) is presented in ref.~\cite{ukk}. The important details of the calculations are: \begin{enumerate} \item A realistic model for the Bethe-Salpeter amplitudes is essential for a realistic estimate of the nuclear effects. We use a recent numerical solution~\cite{uk} of the ladder Bethe-Salpeter equation with a realistic exchange kernel. \item The Bethe-Salpeter amplitudes and, therefore, eqs.~(\ref{f2m})-(\ref{g1m}) have a nontrivial singular structure. These singularities must be carefully taken into account. \item The BS amplitudes are numerically calculated with the help of the Wick rotation. Therefore, the numerical procedure for inverse Wick rotation must be applied. \end{enumerate} \begin{center} \vspace*{-5cm} \let\picnaturalsize=N \def3.0in{4.8in} \defh-1d.eps{g1.ps} \ifx\nopictures Y\else{\ifx\epsfloaded Y\else\input epsf \fi \let\epsfloaded=Y \centerline{\ifx\picnaturalsize N\epsfxsize 3.0in\fi \epsfbox{h-1d.eps}}}\fi \vspace*{-5.5cm} Figure~1. \end{center} Calculations with the realistic BS amplitudes result in the behavior of the $g_1^D$ very similar to other calculations~\cite{kaemp,mg1}. The result is presented in Fig.~1 in the form of the ratio, $g^D_1/g_1^N$. Dotted curve presents non-relativistic calculations with the Bonn wave function, solid curve the BS result. The dashed curve is the illustrative result for the non-relativistic calculations utilizing the BS axial-vector density. Last curve, dot-dased, in Fig.~1 corresponds to the naive formula (\ref{simple}). Despite seemingly drastic difference in the ratio $g^D_1/g_1^N$ given by (\ref{simple}) and realistic calculations, the typical experimental errors today are larger. (Large fluctuations of the ratio at $x<0.7$ are not too important. They correspond to zeros of the nucleon SF which are slightly shifted by the convolution formula.) Indeed, in Fig.~2 we present representative example of data (SMC-1994), together with two fits of these data (dashed lines). The solid lines present results of {\em exact} extraction of the nucleon SF from the present deuteron data. We see that curves for deuteron and nucleon both do not contradict the experiment. However, in certain kinematical conditions effects can be bigger. For instance, lately much interest is devoted to the discussion of the Gerasimov-Drell-Hearn sum rule for the proton and the neutron SFs at small $Q^2$ in general and at $Q^2 = 0$ in particular (see reviews~\cite{jaffer,rev2} and references therein). Very important contribution to the study of the neutron SFs is expected from the Jefferson Lab groups~\cite{expgdh}, where experiments with the deuterons and $^3He$ are planned in the intervals $Q^2\sim 0.15 - 2$~GeV$^2$. Analysis of the deuteron SFs in this interval of $Q^2$, the nucleon ``resonances'' region, shows that effect of the binding and Fermi motion is much larger here than in the deep inelastic regime. An example of the calculation of the deuteron structure function, $g_1^D(x,Q^2)$, is presented in Fig.~3a (dashed line) at $Q^2=1.0$~GeV$^2$. It is compared with the nucleon SF, $g_1^N(x,Q^2)$, input into the calculation. In the areas of resonance structures in $g_1^N(x,Q^2)$, the deuteron SF differs up to 50\%! In Fig.~3b we present a comparison of the neutron SF, $g_1^n$ (solid line, input into calculations in Fig.~3a), with the ``neutron'' SF ``extracted'' by means of the naive formula (\ref{simple}) (dashed line). We see that these two functions have nothing in common. \vspace*{-2mm} \begin{center} \let\picnaturalsize=N \def3.0in{4.0in} \defh-1d.eps{g1-extract.eps} \ifx\nopictures Y\else{\ifx\epsfloaded Y\else\input epsf \fi \let\epsfloaded=Y \centerline{\ifx\picnaturalsize N\epsfxsize 3.0in\fi \epsfbox{h-1d.eps}}}\fi Figure~2. \end{center} The same effects appear in $^3He$~\cite{ciofi}. The presented example, shows that in every particular situation one has to consider the nuclear effects and take into account corresponding corrections to the SFs. Mathematically the problem of extraction of the neutron SF from the deuteron data is formulated as a problem to solve the inhomogeneous integral equation (\ref{g1m}) for the neutron SF with a model kernel and experimentally measured left hand side\footnote{Depending on the model, some additive corrections could be taken into account.}, $g_1^D$. Recently we proposed a method to extract the neutron SF from the deuteron data within any model, giving deuteron SF in the form of a "convolution integral plus/minus additive corrections"~\cite{unfo}. The principal advantages of the method, compared with the smearing factor method, are the following. (i) Only analyticity of the SF need be assumed, (ii) the method allows us to elaborate on the spin-dependent SF, where the traditional smearing factor method does not work. \begin{center} \begin{minipage}{15cm} \vspace*{-7.05cm} \hspace*{-2.5cm} \let\picnaturalsize=N \def3.0in{3.60in} \defh-1d.eps{gdh-sf1.ps} \ifx\nopictures Y\else{\ifx\epsfloaded Y\else\input epsf \fi \let\epsfloaded=Y \centerline{\ifx\picnaturalsize N\epsfxsize 3.0in\fi \epsfbox{h-1d.eps}}}\fi \vspace*{-12.7cm} \hspace*{4.5cm} \let\picnaturalsize=N \def3.0in{3.60in} \defh-1d.eps{gdh-ntr1.ps} \ifx\nopictures Y\else{\ifx\epsfloaded Y\else\input epsf \fi \let\epsfloaded=Y \centerline{\ifx\picnaturalsize N\epsfxsize 3.0in\fi \epsfbox{h-1d.eps}}}\fi \end{minipage} \vskip .9cm a) \hspace*{6cm} b) Figure~3. \end{center} \section{Other spin-dependent structure functions} \subsection{SFs for spin-1 hadron, ${b_{1,2}^D}$}. The SF $b_1^D$ is defined by (see ref.~\cite{ukk,ub} and references therein): \begin{eqnarray} && b_2(x_N,Q^2) = F_2^D(x,Q^2,M=+1)-F_2^D(x,Q^2,M=0),\label{b2} \end{eqnarray} Note, the SF $F_2^D(x,Q^2,M)$ is independent of the lepton polarization, therefore, both SFs, $F_2^D$ and $b_2^D$, can be measured in experiments with an unpolarized lepton beam and polarized deuteron target. In view of eq.~(\ref{pm}), only one of the SFs $F_2^D(x,Q^2,M)$ is needed, in addition to the spin-independent $F_2^D(x,Q^2)$, in order to obtain $b_2(x,Q^2)$. The other SF, $b_1^D$, is related to the deuteron SF $F_1^D$, the same way as $b_2^D$ is related to $F_2^D$, via eqs.~(\ref{f2}), and $b_2^D = 2xb_1^D$. Sum rules for the deuteron SFs $b_1^D$ and $b_2^D$ are a result of the fact that the vector charge and energy of the system are independent of the spin orientation: \begin{eqnarray} \int\limits_0^1 dx_D b_1^D(x_D) =0, \quad \int\limits_0^1 dx_D b_2^D(x_D) =0. \label{sr2} \end{eqnarray} These sum rules were suggested by Efremov and Teryaev~\cite{et}. \vspace*{-2.75cm} \begin{center} \let\picnaturalsize=N \def3.0in{3.90in} \defh-1d.eps{b2.ps} \ifx\nopictures Y\else{\ifx\epsfloaded Y\else\input epsf \fi \let\epsfloaded=Y \centerline{\ifx\picnaturalsize N\epsfxsize 3.0in\fi \epsfbox{h-1d.eps}}}\fi \vspace*{-3.2cm} Figure~4. \end{center} The SFs $b_1^D$ and $b_2^D$ are calculated within two approaches as well. The results are shown in Fig.~4 a) and b). The behavior of the functions in Fig.~4 a) suggests the validity of the first of sum rules (\ref{sr2}). At the same time, the nonrelativistic calculation for $b_2^D$ in Fig.~4 b) (dotted line) obviously does not satisfy the second sum rule. The main difference of the relativistic and nonrelativistic calculations is at small $x$, where these approaches give different signs for the SFs. To check a model dependence of the nonrelativistic calculations, we also performed calculations with the ``softer'' deuteron wave function (with cut-off of the realistic wave function at $|{\bf p}| = 0.7$~GeV). Corresponding SFs are shown in Fig.~4 a) and b) (dashed line). It also does not affect the principle conclusion that the nonrelativistic approach violates the sum rules. \subsection{Chiral-odd SF ${h_{1}^D}$}. The spin-dependent SFs $h_1$ of the nucleons and deuteron can not be measured in the inclusive deep inelastic scattering, but in the semi-inclusive process~\cite{jj}. In this sense these SFs are different from the SFs studied in the present paper. However, we present the results for these functions, since (i) they carry important information about the spin structure of the nucleons~\cite{jj,hj} and the deuteron~\cite{uhk}, (ii) the experiments are planned to measure them~\cite{exp} and (iii) from the theoretical point of view structure function of the deuteron, $h_1^D$, is defined in a way very similar to the usual deep inelastic SFs~\cite{uhk}: \begin{eqnarray} h_1^D(x_N) &=& i \int \frac{d^4p}{(2\pi)^4} {h_1^N}\left( \frac{x_N m}{p_{10}+p_{13}}\right) \label{h1m}\\ &&\frac{\left. {\sf Tr}\left\{ \bar\Psi_M(p_0,{\bf p})\gamma_5\gamma_3\gamma_0 \Psi_M(p_0,{\bf p}) (\hat p_2-m) \right \}\right |_{M=1}}{2(p_{10}+p_{13})}. \nonumber \end{eqnarray} To calculate the realistic SF $h_1^D(x)$ we need the nucleon SFs $h_1^N(x)$. However, so far there is no existing experimental data for this function, and very little is known about the form of $h_1^N$ in theory. In the present paper we follow the ideas of ref.~\cite{jj} to estimate $h_1^N$. Since the sea quarks do not contribute to $h_1^N$, its flavor content is simple: \begin{eqnarray} h_1^N (x) = \delta u(x) +\delta d(x), \label{hn1} \end{eqnarray} where $\delta u(x) $ and $\delta d(x)$ are the contributions of the u- and d-quarks, respectively~\cite{jj,hj}. Since the matrix elements of the operators $\propto \gamma_5 \gamma_3$ and $\propto \gamma_5 \gamma_3\gamma_0$ coincide in the static limit, as a crude estimate we can expect that \begin{eqnarray} \delta u(x) \sim \Delta u(x), \quad \quad \delta d(x) \sim \Delta d(x), \label{dud1} \end{eqnarray} where $\Delta u(x)$ and $\Delta d(x)$ are contributions of the u- and d-quarks to the spin of the nucleon, which is measured through the SF $g_1^N$. Correspondly, the simplest estimation for $h_1^N$ \begin{eqnarray} h_1^N (x) = \alpha \Delta u(x) +\beta \Delta d(x), \label{hn2}\\ \alpha = \beta=1 \label{ab1} \end{eqnarray} should not be too unrealistic. In fact, the bag model calculation shows that difference between $\delta q$ and $\Delta q$ is typically only few percent~\cite{jj}. This analysis is mostly a qualitative one, since it is limited by the case with one quark flavor and does not pretend to describe a phenomenology. To evaluate possible deviations from the simple choice of $h_1^N$, (\ref{hn2}) with (\ref{ab1}), we suggest: \begin{eqnarray} \alpha = \delta u/\Delta u, \quad \beta= \delta d/\Delta d, \label{ab2} \end{eqnarray} where $\delta q$ and $ \Delta q$ are the first moments of $\delta q(x)$ and $ \Delta q(x)$, respectively ($q = u, d$). For $\delta u$ and $ \delta d$ we can adopt the results from the QCD sum rules and the bag model calculations~\cite{hj}. As to $\Delta u$ and $ \Delta d$, we can use the experimental data analysis~\cite{jaffer,rev2} or theoretical results, e.g. the QCD sum rule results~\cite{hj}. Thus, we estimate~\cite{uhk}: \begin{eqnarray} \alpha = 1.5 \pm 0.5, \quad \beta= 0.5 \pm 0.5, \label{ab3} \end{eqnarray} at the scale of $Q^2=1$~GeV$^2$. \vskip -.7cm \begin{center} \let\picnaturalsize=N \def3.0in{3.0in} \defh-1d.eps{h-1d.eps} \ifx\nopictures Y\else{\ifx\epsfloaded Y\else\input epsf \fi \let\epsfloaded=Y \centerline{\ifx\picnaturalsize N\epsfxsize 3.0in\fi \epsfbox{h-1d.eps}}}\fi Figure~5. \end{center} The realistic form of the distributions $\Delta u(x)$ and $\Delta d(x)$ can be taken from a fit to the experimental data for $g_1^N$. In our calculations we used parametrization from ref.~\cite{shaf}. At this point we have to realize that, in spite of expected relations~(\ref{dud1}), distributions $\delta q$ and $\Delta q$ are very different in their nature. Especially at $x {\ \lower-1.2pt\vbox{\hbox{\rlap{$<$}\lower5pt\vbox{\hbox{$\sim$}}}}\ }0.1$, where $\Delta q$ probably contains a singular contribution of the polarized sea quarks, but $\delta q$ does not. Therefore we expect eq.~(\ref{hn2}) to be a reasonable estimate in the region of the valence quarks dominance, say $x {\ \lower-1.2pt\vbox{\hbox{\rlap{$>$}\lower5pt\vbox{\hbox{$\sim$}}}}\ }0.1$. For completely consistent analysis, the parameters $\alpha$ and $\beta$, and the distributions $\Delta u(x)$ and $\Delta d(x)$ should be scaled to the same value of $Q^2$. However, for the sake of the unsophisticated estimates we do not go into such details. The results of calculation of the nucleon and deuteron SFs, $h_1^N$ (solid lines) and $h_1^D$ (dashed lines), are shown in Fig.~5. The group of curves 1 represents case (\ref{ab1}), which is a possible lower limit for $h_1^{N,D}$ in accordance with our estimates (\ref{ab3}). Curves 2 represent the case $\alpha = 1.5,\quad\beta = 0.5$, which is close to the mid point results of the bag model and the QCD sum rules. The upper limit corresponding to the estimates (\ref{ab3}) is presented by curves 3. For all cases the deuteron SF is suppressed comparing to the nucleon one, mainly because of the depolarization effect of the D-wave in the deuteron. This is quite similar to the case of the SFs $g_1^N$ and $g_1^D$. Note that our estimate of the nucleon SF $h_1^N$, (\ref{ab3}), gives systematically larger function than naive suggestion (\ref{ab1}), the curves 1 in Fig.~5 which essentially corresponds to the estimate $h_1^N \simeq (18/5) g_1^N$, neglecting possible negative contribution of the s-quark sea~\cite{jaffer,rev2}. The large size of the effect suggests that it can be detected in future experiments with the deuterons~\cite{exp}. \section{Brief conclusion} We have presented the results of our study of the spin-dependent structure functions of the deuteron. In particular, the leading twist $g_1^D$, $b_{1,2}^D$ and $h_1^D$ are considered. The issue of the extraction of the neutron structure functions from the deuteron data is addressed. The role of relativistic effects is studied and can be summarized as: (i) relativistic calculations give a slightly larger magnitude of the binding effects, (ii) the relativistic Fermi motion results in ``harder'' SF at high $x$, and (iii) covariant approach is internally consistent, while the nonrelativistic approach is internally inconsistent and violates important sum rules. \section*{Acknowledgements} We wish to thank every one who essentially contributed to studies included in this presentation, L.P. Kaptari, C. Ciofi degli Atti, Han-xin He and S. Scopetta. This work is supported in part by NSERC, Canada, and INFN, Italy. {\small
proofpile-arXiv_065-454
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{INTRODUCTION} \noindent The occurrence of variable stars radially pulsating in the fundamental and/or the first overtone modes is a well known and well established observational evidence based on both Population I and Population II variables. The possible occurrence of stars pulsating in the second overtone (SO) is a still much debated argument. For a long time the observational scenario concerning such an occurrence has been limited to the suggestions of several authors (van Albada \& Baker 1973; Demers \& Wehlau 1977; Clement, Dickens \& Bingham 1979; Nemec, Wehlau \& Mendes de Oliveira 1988 and references therein) that some globular clusters RRc variables (first overtone) characterized by very short periods (P$\sim$ 0.3 d) and small pulsational amplitudes could be good SO~ candidates. Only in recent times the large and homogeneous photometric databases collected by the MACHO project for radial pulsators in the Large Magellanic Cloud (LMC) have brought out for the first time a sound evidence of the occurrence of SO pulsators in classical Cepheids (Alcock et al. 1995). On the basis of the same extensive photometric survey, from the period distribution of about 8000 RR Lyrae variables, it has been also suggested (Alcock et al. 1996) that a peculiar peak located at P=0.28 d could be due to low mass SO~ variables. \noindent From a theoretical standpoint, the limit cycle stability of second overtone~ pulsators in population I and population II variable stars has been an odd problem for a long time. As a matter of fact, still at present we lack either firm theoretical predictions concerning the existence of SO~ pulsators, or convincing insights on the physical mechanisms which could govern their approach to mode stability. Classical nonlinear analyses of pulsation properties and modal stability of population II low mass variables (Christy 1966; Stellingwerf 1975) never produced any firm evidence of unstable SO~ pulsators. By using a simple nonlinear and nonadiabatic one-zone model Stellingwerf, Gautschy \& Dickens (1987, hereinafter referred to as SGD) succeeded in producing an unstable SO~ pulsator. However, even though one-zone pulsating models can be of some help to disclose the main processes which constrain the physical structure of stellar envelopes, unfortunately this approach cannot supply definitive conclusions on the existence of these variables since these models do not take into account the dynamical effects of the regions which operate the damping of the pulsation. An unstable SO model at limiting amplitude was constructed for RR variables by Stothers (1987) on the basis of a nonlinear radiative approach. However, the light curve of his model shows a peculiar feature, i.e. a deep splitting just after the maximum in luminosity, which is not observed in any known group of radial pulsators and in the light curve predicted by SGD. Moreover, no evidence of unstable SOs was found in the homogeneous and systematic survey of nonlinear, nonlocal and time-dependent convective models provided by Bono \& Stellingwerf (1994 hereinafter referred to as BS) and by Bono et al. (1996a hereinafter referred to as BCCM). \noindent According to the previous discussion, no SO~ pulsators appear to be foreseen in old, low mass variable stars. In this context, it is worth underlining that periods of RRc in globular clusters are not in conflict with first overtone expectations, and their small amplitudes could be due to the sudden decrease of the first overtone amplitudes near the first overtone blue boundary (BCCM). Moreover, the peak in the period distribution at P=0.28 d of RR Lyrae in LMC can be taken as a compelling evidence of a moderately metal rich population, with periods which overlap the characteristic short periods already found both in the Galactic bulge and in the metal rich field RR Lyrae. The reader interested to a detailed discussion on the evolutionary and pulsational properties of metal rich RR Lyrae variables is referred to Bono et al. (1996b). \noindent As far as more massive stars are concerned, it has been often suggested that the Hertzsprung progression of the bump in the light curve of classical Cepheids could be closely correlated to a {\em spatial resonance} between the SO~ and the fundamental mode (Simon \& Schmidt 1976; Buchler et al. 1996). However, present nonlinear calculations devoted to the modal stability of these variables did not supply any firm conclusion on the existence of SO~ pulsators. Although further nonlinear investigations have not been so far undertaken to properly tackle the problem, it is worth mentioning that linear nonadiabatic models provided by Baker (1965) for the mass value $M/M_{\odot}\,$=1.0 disclosed a SO~ instability over a wide range of luminosities and temperatures. Moreover, Deupree \& Hodson (1977), in their study on the instability strip of Anomalous Cepheids, gave an evaluation of the SO~ linear blue edge for even larger masses ($1.0 \leq $ $M/M_{\odot}\,$ $\leq 2.0$). However, the quoted authors did not come to a sound conclusion since the numerical difficulties caused by the increased efficiency of the convective overshooting prevented them from evaluating the nonlinear modal stability of these pulsators. \noindent An extensive grid of $\delta$ Scuti~ pulsation properties has been recently provided by Milligan \& Carson (1992) taking into account a combination of stellar evolution models and both linear and nonlinear pulsation models. Although in this investigation they carried out a detailed analysis of the linear and nonlinear properties of $\delta$ Scuti~ stars, only few models have been followed at limiting amplitude due to the small value of the growth rates involved. Therefore the ultimate modal stability of the wide range of nonlinear models could not be firmly established. Let us note that $\delta$ Scuti~ variables are most attractive objects among the several groups of variables presently known since they also show simultaneous excitation of both radial and nonradial modes. This peculiar feature is of the utmost importance for properly addressing several astrophysical questions concerning the physical mechanisms which rule the driving and the quenching of stellar oscillations. \noindent The observational panorama of $\delta$ Scuti~ stars is rather complex since both the rapidly oscillating roAp magnetic stars (Shibahashi 1987; Martinez \& Kurtz 1995) and the large amplitude metal poor SX Phoenicis stars (Nemec, Linnell Nemec \& Lutz 1994; Nemec et al. 1995) belong to this group of variable stars as well. However, new high quality photometric data of $\delta$ Scuti~ variables in the Galactic field (Rodriguez et al. 1995), in the Baade's window (Udalski et al. 1995), in open clusters (Frandsen et al. 1995) and in globular clusters (Nemec et al. 1994) have been recently provided and therefore a proper classification of these objects based on their evolutionary and pulsational properties will be soon available. Another interesting aspect which makes $\delta$ Scuti~ variables worth being investigated is that during their evolution intermediate mass stars cross the instability strip in different evolutionary phases. In fact, in contrast with canonical Cepheid-like variables which are connected with Helium burning evolutionary phases, in this region of the HR diagram the current evolutionary theory foresees stars in pre-main-sequence, main sequence and post main sequence phases. As a consequence, their pulsational properties can supply useful constraints on their evolutionary properties and at the same time an independent check of stellar models (Breger 1993,1995). \noindent Linear adiabatic and nonadiabatic models of these objects have been computed by Stellingwerf (1978), Andreasen, Hejlesen \& Petersen (1983) and Andreasen (1983). Even though these investigations have significantly improved our knowledge of the location of the linear blue boundaries of the instability strip, the theoretical scenario of $\delta$ Scuti~ stars presents several unsettled problems. According to this evidence, we decided to extend the investigation on modal stability and pulsational properties already provided for low mass He burning stars to larger mass values. Canonical evolutionary prescriptions indicate that stars with masses of the order of (1.5 - 2.0)$M_{\odot}$, unlike low mass stars, spend a non negligible portion of their lifetime inside the instability strip during their evolution off the Main Sequence. Therefore relatively massive stars are expected to generate, during this phase, short period radial pulsators. \noindent In this paper we investigate the nonlinear pulsational properties of this kind of variable stars (Eggen 1994; McNamara 1995) by exploring the pulsational behavior of a sequence of models as a function of the surface effective temperature. For the first time we show that the adopted theoretical framework accounts for the modal stability of SO~ pulsators and allows some predictions concerning the full amplitude behavior of SO~ light and velocity curves. In \S 2 we describe the numerical and physical assumptions adopted for constructing both linear and nonlinear pulsation models, whereas in \S 3 we discuss the full amplitude pulsation behavior for the first three modes. The shape and the secondary features of both light and velocity curves are presented in \S 4 together with a plain comparison with available photometric data. The physical parameter which rules the SO~ limit cycle stability and their dynamical properties are discussed in \S 5. Our conclusions are presented in \S 6. In this section the consequences arising from these new findings and the observative features worth being investigated are outlined as well. \section{STANDARD PULSATION MODELS} \noindent The theoretical framework adopted for investigating linear and nonlinear pulsation characteristics of radial pulsators have been previously described in a series of papers (Stellingwerf 1982; BS and references therein). The sequence of models presented in this paper was computed at fixed mass value ($M/M_{\odot}\,$ =2.0), chemical composition (Y=0.28, Z=0.02) and luminosity level (log $L/L_{\odot}\,$ =1.7) and by exploring a wide range of effective temperatures ($7500 \geq T_e \geq 6000$ K). The static envelope models were constructed by adopting an optical depth of the outermost zone of the order of 0.001, and the inner boundary was assumed fixed so that possible destabilization effects due to variations in the efficiency of the H shell burning are ignored. Moreover, we adopted the OP radiative opacities provided by Seaton et al. (1994) for temperatures higher than 10,000 K whereas for lower temperatures we adopted molecular opacities provided by Alexander \& Ferguson (1994). The method adopted for handling the opacity tables was already described in Bono, Incerpi \& Marconi (1996). \noindent Each envelope model extends from the surface to 20-10\% of the stellar radius and the zone closest to the Hydrogen Ionization Region (HIR) was constrained to $T_{HIR}=1.3 \times 10^4$ K. Between the HIR and the surface we inserted 20 zones to ensure a good spatial resolution of the outermost regions throughout the pulsation cycle. Due to the key role played by spatial resolution for firmly estimating both linear and nonlinear modal stability of higher modes, the envelopes were discretized by adopting a detailed zoning in mass. The mass ratio between consecutive zones ($\theta$) has been assumed equal to $\theta$=1.1 for temperatures lower than $6 \times 10^5$ K, whereas for higher temperatures it has been set equal to $\theta$=1.2. By adopting this type of zoning the ionization regions and the opacity bump due to iron are covered with a number of zones lying between 100 and 150. The fine spatial resolution of the ionization fronts provides, in turn, an accurate treatment of both the formation and the propagation of shock fronts during the phase interval between the phase of minimum radius and the phase of maximum luminosity (for complete details see BS). On the basis of these assumptions a typical envelope model is characterized by roughly 20-30\% of the total stellar mass and by 150-250 zones. \noindent To clarify matters concerning the dependence of the pulsation behavior on the spatial resolution we have constructed a much finer envelope model. This model is located at $T_e=7000$ K and, in contrast with the standard sequence of linear nonadiabatic models, it was constructed by adopting a smaller mass ratio ($\theta$=1.08) for the regions located at temperatures lower than $6 \times 10^5$ K. Moreover, the inner boundary condition for this model was chosen in such a way that the base of the envelope was located below 0.1 of the photospheric radius and its temperature was of the order of $5-6 \times 10^6$ K. \noindent As a consequence of these assumptions the detailed model presents an increase in both the envelope mass ($M_{env}=0.46 M_T$ against $M_{env}=0.21 M_T$) and the number of zones which cover the ionization regions (180 against 150). We eventually found that for the first three modes the differences between the pulsation characteristics of both detailed and standard model are quite negligible. In fact, the discrepancies range from $10^{-4}$ to $10^{-3}$ for the periods and from $10^{-6}$ to $10^{-4}$ for the growth rates. As a result we can assume that the spatial resolution of the standard sequence allows a proper treatment of the radial pulsation of $\delta$ Scuti~ stars during their off main sequence evolution. \section{APPROACH TO LIMIT CYCLE STABILITY} \noindent According to the usual approach, a sequence of linear nonadiabatic models was first constructed for supplying the static structure of the envelope to the nonlinear stability analysis. Then the equations governing both the dynamical and the convective structures were integrated in time until the initial perturbations and the nonlinear fluctuations, which result from superposition of higher order modes, settled down (for more detail see Bono, Castellani \& Stellingwerf 1995). The dynamical behavior of the envelope models was computed for the first three modes and the initial velocity profile was obtained by perturbing the linear radial eigenfunctions with a constant velocity amplitude of 20 km$s^{-1}$ which causes a global expansion of the envelope. \noindent As is well known, the linear nonadiabatic $\delta$ Scuti~ models present very small growth rates and therefore, before radial motions approach the nonlinear limit cycle stability, it is necessary to carry out extensive calculations. In fact, the long-term stability of a particular mode, due to the mixture of both periodic and nonperiodic motions characterized by very small growing and/or decaying rates, cannot be easily assessed at small amplitude. As a consequence, in order to find out the possible appearance of a mode switching or of a mixture of modes we evaluated the asymptotic behavior of each mode by performing very long runs. This approach leads to an integration of the governing equations for a number of periods which ranges from 5,000 to 50,000 for some peculiar cases. The integration is generally stopped when the nonlinear total work is vanishing and the pulsation amplitudes present over two consecutive periods a periodic similarity of the order of or lower than $10^{-(2 \div 3)}$. \noindent Since this is the first time that hydrodynamic calculations are performed over such a long time interval, Figures 1, 2 and 3 show the time behavior of period, velocity and magnitude for three cases characterized by a different approach to nonlinear limit cycle stability. In particular, Fig. 1 shows the variation of the quoted quantities for a single pure SO~ model, whereas Fig. 2 shows that radial motions at $t\approx 7$ yrs experience a mode switching from the first overtone to the SO~. Figure 3 finally presents the limiting amplitude behavior of a case which presents a permanent mixture of different radial modes. \noindent As a result of the modal stability analysis, we found stable nonlinear limit cycles in the fundamental, in the first overtone and, for the first time, in the second overtone when the effective temperature is increased. Even though so far the location inside the instability strip of $\delta$ Scuti~ stars characterized by different pulsation modes has not been firmly established, the previous finding confirms the distribution originally suggested by Breger \& Bregman (1975). In fact, by assuming that $\delta$ Scuti~ variables are radial pulsators, these authors found that observed second and first overtone variables were located at effective temperatures higher than the fundamental ones. \noindent In Table 1 are listed selected observational parameters for the sequence of $\delta$ Scuti~ models. As a first result, data in Table 1 show that theoretical periods appear in general agreement with the observed range of $\delta$ Scuti~ values. It is worth underlining that the effective temperature of the fundamental red edge should be considered an upper limit. As a matter of fact, even though the model located at 6300 K after 23,000 periods presents both a constant negative value in the total work term and a very low pulsational amplitude ($\Delta M_{bol} \approx 4-6 \times 10^{-3}$ mag), this region of the instability strip, due to the slow approach to limit cycle stability, should be investigated in more detail before firmly constraining the location of the red edge. \noindent In order to disclose the main features of the modal behavior in $\delta$ Scuti~ stars, Figure 4 shows the bolometric light curves and the surface radial velocities of SO~ pulsators; Figure 5 shows the same quantities but they are referred to selected first overtone (solid lines) and fundamental (dashed lines) pulsators. Figure 6 shows the light and radial velocity curves of mixed mode pulsators, i.e. of models which present a permanent mixture of different radial modes at limiting amplitude. \noindent Inspection of light curves discloses the surprising evidence that the shape of SO~ light curves -sudden increase in the rising branch and slow decrease in the decreasing branch- closely resembles canonical fundamental mode rather than first overtone RR Lyrae pulsators. This finding confirms the original prediction concerning the shape of SO~ light curves made by SGD. Moreover, we find that moving from SO~ to lower pulsational modes the amplitudes progressively decrease and the shape of the light curves becomes more sinusoidal. It is worth noting that the pulsation amplitudes of RR Lyrae variables, which are located in the same region of the instability strip, present an opposite trend. In fact, for this group of pulsators the RRab variables (fundamental) show the largest amplitudes. These theoretical prescriptions can be usefully compared with the observational scenario recently discussed by McNamara (1995 and references therein). \noindent According to this author, $\delta$ Scuti~ stars on the basis of their luminosity amplitude can be empirically divided into two groups. The light curves of stars with larger amplitudes appear to be asymmetrical whereas the light curves for lower amplitudes tend to be much more symmetrical. However, in the above paper McNamara also suggests that for low amplitude variables, which are poorly sampled, it is often difficult to determine whether the light curves are symmetric or asymmetric. Therefore, for light curves which are only partially covered by photometric data, several cases of probable asymmetry are brought forward. By analogy with the behavior of RR Lyrae variables, McNamara (1995) assumes that stars with asymmetric, large amplitude light curves are fundamental pulsators whereas symmetric, low amplitude light curves belong to first overtone pulsators. The comparison of similar empirical prescriptions with the current theoretical scenario shows a convincing degree of agreement. However, theory now tells us that asymmetric, large amplitude pulsators are good SO~ candidates, whereas low amplitude pulsators could be a mixture of fundamental and first overtones. \noindent To go further on with this comparison, let us refer to the sample of variable stars recently collected by the OGLE collaboration (Udalski et al. 1995 and references therein) as the result of their search for evidence of microlensing in the bulge of the Galaxy. Inspection of photometric data connected with short period variables discloses that the observed light curves can be arranged in three typical classes, as shown in Fig. 7, with class "A" representing McNamara large amplitude pulsators and classes "B" and "C" the small amplitudes ones. For a meaningful comparison between theory and observation, the bolometric light curves have been transformed into the I band according to Kurucz's (1992) atmosphere models. Figures 8 and 9 show the light curves for single mode pulsators. Due to the magnitude scale adopted for plotting data in Fig. 9, the light curves of fundamental pulsators (dashed lines) seem almost perfectly sinusoidal. However, even though the luminosity variations throughout the cycle are quite smooth, a bump appears before the phase of minimum radius. Taking into account that we explored only one mass value and only one luminosity level, the comparison should be considered more than satisfactory. \section{ LIGHT CURVE MORPHOLOGY} \noindent Available observational data hardly allow to detect minor details in the light curves. However, the quality of both spectroscopic and photometric data is rapidly improving (see for example Milone, Wilson \& Fry 1994; Breger et al. 1995 and references therein) and therefore in this section we discuss even minor features of theoretical light curves in order to underline the theoretical predictions worth being investigated with the required accuracy. As a first point, let us notice that the light and velocity curves of SO~ pulsators present two further relevant distinctive features: \noindent 1) like canonical first overtone RR Lyrae variables, the bump does not appear along the decreasing branch of the light curves and, moving from higher to lower effective temperatures the dip becomes more and more evident along the rising branch; \noindent 2) moving from the blue to the red boundary of the SO instability region the velocity curves show smooth variations but at phases 0.2-0.3 a bump appears due to the propagation of an outgoing shock. \noindent As a second point, we find that the shape of first overtone light curves presents some features which allow a careful distinction between different radial modes. In fact for these models the dip is the main maximum, whereas the "true" maximum takes place along the decreasing branch. Moreover, the first overtone light curves show that the bump appears along the increasing branch and that just before the phase of minimum radius they also display a short stillstand phase ($\phi \approx 0.45$). \noindent The scenario concerning the pulsation characteristics of $\delta$ Scuti~ stars can be now nicely completed by the models which are simultaneously excited in two or more radial modes. Figure 6 shows a collection of light and velocity curves of mixed mode pulsators, a glance to these curves brings out both the expected fluctuation of the pulsation amplitudes between consecutive cycles and the appearance of the secondary features which characterize the first three lower single mode pulsators. \noindent A thorough comparison with observed period ratios of $\delta$ Scuti~ variables is beyond the scope of the present study since the "true" period ratios should be evaluated through a Fourier decomposition of the theoretical light curves. Moreover, for properly constraining the stellar masses and the luminosity levels of these objects by means of the Petersen diagram ($P_1$/$P_0$ vs. $P_0$) an extensive set of nonlinear models computed for different assumptions on astrophysical parameters is necessary (Bono et al. 1996c). \noindent Nevertheless, since period ratios of double mode pulsators can provide valuable clues on several astrophysical problems involving both the evolutionary and the pulsational properties of these objects, we constructed a new sequence of detailed, linear, radiative, nonadiabatic models by adopting the assumptions already discussed at the end of section 2. This analysis has been undertaken only for supplying a preliminary but meaningful theoretical guess concerning the location of this group of variables inside the instability strip. \noindent At first it is important to note that the linear periods and the related period ratios listed in Table 2 are, within the estimated uncertainties, in agreement with observed values (see for example data on double mode $\delta$ Scuti~ stars collected by Andreasen 1983 and by Petersen 1990). Indeed the observed period ratio between first overtone and fundamental pulsators ranges, for large amplitude $\delta$ Scuti~ stars, from 0.760 to 0.780, whereas the period ratio between second overtone and first overtone is approximatively of the order of 0.800. This agreement is a remarkable result since the period ratios predicted by pulsation models are the most important observable adopted for finding out whether the pulsation is "driven" by radial or by nonradial modes. As a consequence, this finding provides sound evidence that these variables are mixed mode radial pulsators (Breger 1979). \noindent Moreover, previous linear and nonlinear results suggest that, due to the appearance of three different modal stabilities inside the instability strip, double mode pulsators belonging to this group of variables are located close to the fundamental blue edge when they are a mixture of fundamental and first overtone and close to the first overtone blue edge when they are a mixture of first and second overtones. A detailed investigation on the dependence of this peculiar occurrence among radial pulsators together with a straightforward analysis of the envelope structure will be discussed in a forthcoming paper (Bono et al. 1996d). \section {SECOND OVERTONE INSTABILITY} \noindent In order to properly identify the regions of the stellar envelope which {\em drive} or {\em damp} the pulsation instability of SO~ pulsators, Fig. 10 shows the nonlinear differential work integrals versus the logarithm of the external mass for a model located close to the SO~ blue edge. In this plane the positive areas denote driving regions (growing oscillations) whereas the negative areas damping regions (quenching oscillations). The total work curve shows quite clearly the two driving sources due to the hydrogen and helium ionization zones as well as the radiative damping due to the inner regions. Unlike in canonical cluster variables, the second helium ionization zone provides a stronger destabilization if compared with the HIR. This effect is mainly due to the increase in effective temperature which causes a shift of the HIR toward the surface and therefore a decrease of the mass which lies above these layers. However, the total work plotted in Fig. 10 clearly shows that this element, in contrast with previous qualitative arguments, provides a substantial amount of driving to the pulsation instability of $\delta$ Scuti~ stars. For the reasons previously discussed all other nonlinear work terms supply a negligible damping effect on the pulsation. \noindent Nevertheless the physical parameter which rules the SO~ instability is the location of the nodes \footnote{In the case of a vibrating membrane a {\em node} is the point where an eigenfunction vanishes or attains a sharp minimum and its phase changes by almost $\pi$ radians.} inside the envelope. In fact, the nodes of temperature, luminosity and radius fall within the region of radiative damping. As a consequence, the amount of damping is strongly reduced and the destabilization of the ionization regions pumps up the pulsation amplitude. In particular, it is worth emphasizing that among cluster variables the outermost node of radius is located quite close to the helium driving region (Stellingwerf 1990). In this context the node of radius has an opposite effect since it reduces the amount of driving and consequently quenches the oscillations. \noindent On the basis of a well-known theorem concerning the Sturm-Liouville eigenvalue problem, a SO~ eigenfunction should split its domain into two parts by means of its {\em nodal lines} (Courant \& Hilbert 1989). Even though during a full pulsation cycle the radial motions never exactly approach the initial equilibrium configuration, the velocity curves plotted in Fig. 11 undoubtedly show two different subdomains characterized by opposite radial motions. In order to disclose the dynamical structure of the model previously discussed, Fig. 11 shows the radial displacements of the whole envelope over two consecutive periods. Dots and pluses mark the phases during which each zone is contracting or expanding respectively. It is easy to ascertain from the above figure that at a fixed pulsation phase the envelope is divided into two different regions and the velocity curves are exactly 180$^{\circ}$ out of phase at their boundaries. In particular, the layers located close to the two boundaries define two zones which virtually remain at rest throughout the pulsation cycle and in which the radial velocity changes its sign abruptly. \section{CONCLUSIONS} \noindent In this paper we report the first theoretical evidence for the occurrence of pulsators that when moving inside the instability strip from lower to higher effective temperatures show three different stable pulsation modes, namely fundamental, first and second overtone. We found that the predicted features of the light curves appear in general agreement with observational constraints concerning $\delta$ Scuti~ variables, suggesting that SO~ pulsators have been already observed and pointing out, at the same time, a revised approach to observational evidences for different pulsation modes. Moreover, we show that the location of radius, luminosity and temperature nodes in the damping region of the stellar envelope is the main physical parameter which governs the limit cycle stability of SO~ pulsators. This result strongly supports the discussion given in the introduction to this paper about the absence of SO~ pulsators in low mass variable stars. \noindent The satisfactory agreement between the computed period ratios and the observed ones casts new light on the problem of limit cycle stability of mixed mode variables belonging to population II stars and provides a valuable piece of information for accounting for the pulsation properties of $\delta$ Scuti~ stars. At the same time there is a growing strong evidence that these stars during their evolution off the main sequence are pure or mixed mode radial pulsators. \noindent Before being able to provide a sound comparison with available photometric data on $\delta$ Scuti~ variables, the present sequence of nonlinear models should be extended to both lower luminosity levels and to smaller mass values. However, even though the computed models cover a restricted range of effective temperatures, the theoretical framework currently adopted accounts for both pure and mixed mode radial pulsators. This new scenario presents several interesting features since both the modal stability and the pulsational behavior have been investigated in a homogeneous physical context without invoking unpleasant {\em ad hoc} physical mechanisms and/or peculiar characteristics. \noindent Plenty new high quality photometric data on $\delta$ Scuti~ stars will be soon available as a by-product of the international projects involved in the search for microlensing events and therefore new sequences of nonlinear $\delta$ Scuti~ models at full amplitude, albeit such analysis places nontrivial computational efforts, are necessary to firmly accomplish the pulsation properties of these objects. \noindent Finally we suggest that short period RR Lyrae-like pulsators found in the Galactic bulge as well as in dwarf spheroidals like Carina and Sagittarius should be regarded as an evidence of relatively massive stars, a witness of the efficiency of star formation until relatively recent times. On the other hand this finding stresses once again the key role played by variable stars as tracers of stellar populations which experienced different dynamical and/or chemical evolutions. \noindent It is a pleasure to thank F. Pasian and R. Smareglia for their kind and useful sharing of computing facilities which made this investigation possible. We are grateful to J. Nemec for his clarifying suggestions as referee on an early draft of this paper. We also benefit from the use of the SIMBAD data retrieval system (Astronomical Data Center, Strasbourg, France). This work was partially supported by MURST, CNR-GNA and ASI. \pagebreak
proofpile-arXiv_065-455
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\subsection*{Acknowledgments} This work was supported by the US Department of Energy, Nuclear Physics Division, under contract number W-31-109-ENG-38.
proofpile-arXiv_065-456
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} \label{intro-sec} Japanese is a language well-known for grammaticization of discourse function. It is rich with ways for speakers to indicate the information status of the discourse entities they are talking about. Japanese allows a speaker to clearly indicate topic-hood, along with the grammatical functions such as subject, object and object2, by using the morphological case markers {\em wa, ga, o, ni}. In addition, it provides morphological means to indicate speaker's perspective through the use of verbal compounding, i.e. the addition of suffixes such as {\em kureta, kita} (See section 3). Unexpressed arguments of the verb are common; these are known as zero pronouns. Because there are zero pronouns and because Japanese is a head-final language with otherwise relatively free word order, there could, in principle, be a great deal of ambiguity. However this is not the case. Speakers are assumed to be cooperative, to be collaborating with the hearer in conversation, and to be ensuring that each utterance is relevant and coherent in the context of what was said before \cite{Grice75,SSJ74}. We believe that speakers do not choose to express their thoughts through arbitrary syntactic constructions, but that there is some correspondence between choice of syntactic construction, what the speaker wants to convey, and aspects of the current discourse situation\cite{Prince85}. Within a theory of discourse, {\sc centering} is a computational model of the process by which a speaker and hearer make obvious to one another their assumptions about the salience of discourse entities. Using pronominal referring expressions is one way for discourse participants to do this. We propose that the resolution of zero pronouns is constrained by centering, and ambiguity is thereby reduced. Centering has its computational foundations in the work of Grosz and Sidner\cite{Grosz77,Sidner79,GS85} and was further developed by Grosz, Joshi and Weinstein\cite{GJW83,GJW86,JW81}. It is formalized as a system of constraints and rules, which can, as part of a computational discourse model, act to control inferencing\cite{JW81}. Brennan, Friedman and Pollard use these rules and constraints to develop an algorithm for resolving the co-specifiers of pronouns\cite{BFP87,Walker89b}. Our analysis uses an adaptation of this algorithm. By making full use of the centering formalism, we avoid the postulation of additional mechanisms, e.g. property sharing\cite{Kameyama86a}. In addition, we propose a notion of {\sc topic ambiguity}, which characterizes some ambiguities in Japanese discourse that are allowed by the centering process. Topic ambiguity has been ignored in previous accounts of Japanese zero pronoun resolution, but it explains the availability of interpretations that previous accounts would predict as ungrammatical. Centering gives us a computational way of determining when a zero pronoun may be assigned {\sc Topic}. This analysis informs the design of language independent discourse processing modules for Natural Language systems. We propose that the centering component of a discourse processing module can be constructed in a language independent fashion, up to the declaration of a language-specific value for one variable in the algorithm, i.e., Cf list ranking (see section \ref{cent-form}). The centering algorithm has been implemented in an HPSG Natural Language system with both English and Japanese grammars. \section{The Centering Formalism} \label{cent-form} The modeling of attentional state in discourse by centering depends on analyzing each pair of utterances in a discourse according to a set of transitions. These transitions are a measure of the coherence of the segment of discourse in which the utterance occurs. Each utterance in a discourse has associated with it a set of discourse entities called {\sc forward-looking centers}, $\rm{Cf}$, and a special member of this set called the {\sc backward-looking center}, $\rm{Cb}$. The {\sc forward-looking centers} are ranked according to discourse salience; the highest ranked member of the set is the {\sc preferred center}, $\rm{Cp}$. With these definitions we can give the constraints: \nopagebreak \begin{itemize} \item {\bf CONSTRAINTS} \\ For each $\rm{U_{i}}$ in a discourse segment $\rm{U_{1}, \ldots ,U_{m}}$: \vspace{-3.0ex} \begin{enumerate} \item There is precisely one $\rm{Cb}$. \item Every element of $\rm{Cf(U_{i}})$ must be realized\footnote {An utterance U (of some phrase, not necessarily a full clause), {\em realizes\/} {\bf c} if {\bf c} is an element of the situation described by U, or {\bf c} is the semantic interpretation of some subpart of U.} in $\rm{U_{i}}$. \item The center, $\rm{Cb(U_{i})}$, is the highest-ranked element of $\rm{Cf(U_{i-1})}$ that is realized in $\rm{U_{i}}$. \end{enumerate} \end{itemize} The typology of transitions from one utterance, $\rm{U_{i}}$, to the next is based on two factors: whether the backward-looking center, $\rm{Cb}$, is the same from $\rm{U_{i-1}}$ to $\rm{U_{i}}$, and whether this discourse entity is the same as the preferred center, $\rm{Cp}$ of $\rm{U_{i}}$. Backward-looking centers are often pronominalized and discourses that continue centering the same entity are more coherent than those that shift from one center to another. This means that some transitions are preferred over others. These two facts give us the rules: \begin{itemize} \item {\bf RULES} \\ For each $\rm{U_{i}}$ in a discourse segment $\rm{U_{1}, \ldots ,U_{m}}$: \vspace{-3.0ex} \nopagebreak \begin{enumerate} \item If some element of $\rm{Cf(U_{i-1})}$ is realized as a pronoun in $\rm{U_{i}}$, then so is $\rm{Cb(U_{i})}$. \item Transition states are ordered. {\sc Continuing\/} is preferred to {\sc retaining\/} is preferred to {\sc shifting-1} is preferred to {\sc shifting}\footnote{\cite{BFP87} introduces the distinction between SHIFTING-1 and SHIFTING.}. \end{enumerate} \end{itemize} The transition states that are used in the rules are defined in Figure \ref{state-fig}, ({\sc backward-looking center} $ = \rm{Cb}$, {\sc preferred Center} $ = \rm{Cp}$). \begin{small} \begin{figure}[htb] \setlength{\unitlength}{.65in} \begin{flushright} \begin{picture}(4,2) \put(0,0){\framebox(2,1){RETAINING}} \put(0,1){\framebox(2,1){CONTINUING}} \put(2,1){\framebox(2,1){SHIFTING-1}} \put(2,0){\framebox(2,1){SHIFTING}} \put(1,2.2){\makebox(0,0){$\rm{Cb(U_{i}) = Cb(U_{i-1}}$)}} \put(3,2.2){\makebox(0,0){$\rm{Cb(U_{i}) \neq Cb(U_{i-1}}$)}} \put(-.33,1.65){\makebox(0,0){$\rm{Cb(U_{i})}$}} \put(-.25,1.45){\makebox(0,0){$=$}} \put(-.33,1.25){\makebox(0,0){$\rm{Cp(U_{i})}$}} \put(-.33,0.70){\makebox(0,0){$\rm{Cb(U_{i})}$}} \put(-.25,0.5){\makebox(0,0){$\neq$}} \put(-.33,0.30 ){\makebox(0,0){$\rm{Cp(U_{i})}$}} \end{picture} \end{flushright} \normalsize \caption{ Transition States} \label{state-fig} \end{figure} \end{small} The centering algorithm incorporates these rules and constraints in addition to linguistic constraints on coreference\cite{BFP87}. The behavior of the centering algorithm for the resolution of pronouns is largely determined by the ranking of the items on the forward center list, $\rm{Cf}$, because, as per Constraint 3, this ranking determines from among the elements that are realized in the next utterance, which of them will be the $\rm{Cb}$ for that utterance. Although all of the factors that contribute to the $\rm{Cf}$ ranking have not been determined, syntax and lexical semantics have an effect\cite{Prince81,Prince85,Hudson88,Brennan89,GJW86,JW81,BF83}. We postulate that this ordering will vary from language to language depending on the means the language provides for expressing discourse functions. Our adaptation of the algorithm for Japanese consists of substituting a different ranking of the forward centers list $\rm{Cf}$. In every other way, the algorithm functions exactly as it is for English. \section{Centering in Japanese} In order to apply the centering algorithm to the resolution of zero pronouns in Japanese, we must determine how to order the forward centers list, $\rm{Cf}$. The function {\sc topic} is indicated by the morphological marker {\em wa}, along with {\sc subject} ({\em ga}), {\sc object} ({\em o}), and {\sc object2} ({\em ni}). The optional use of {\em wa} picks out the most salient entity in the discourse. In addition, Kuno proposed the notion of {\sc empathy}, which is the perspective from which a speaker describes an event\cite{Kuno73}. The realization of speaker's empathy is especially important when describing an event involving some transfer. For example, there is no way to describe a {\em giving\/} and {\em receiving\/} situation objectively\cite{Kuno-Kab77}. In (1), the use of the past tense {\em kureta} of the verb {\em kureru\/}, indicates the speaker's empathy with the discourse entity realized in object position\footnote{We use identifiers of all capital letters to denote the discourse entity realized by the corresponding string. Centers are semantic entities, not syntactic ones.}. \noindent (1)\\ \footnotesize \begin{tabular}{llllll} Hanako wa & Taroo ni & hon o & kureta. & & \\ top-subj & obj2 & book obj & give-past & & \\ \multicolumn{6}{l}{{\it ``Hanako gave Taroo a book.''}}\\ \multicolumn{6}{l}{EMPATHY=OBJ2=TAROO} \\ \end{tabular} \normalsize In (2), the speaker's empathy with the subject entity's perspective is indicated using {\em yatta}, the past tense of the verb {\em yaru\/}. \noindent (2)\\ \footnotesize \begin{tabular}{llllll} Hanako wa & Taroo ni & hon o & yatta. & & \\ top-subj & obj2 & book obj & give-past & & \\ \multicolumn{6}{l}{{\it ``Hanako gave Taroo a book.''}}\\ \multicolumn{6}{l}{EMPATHY=SUBJ=HANAKO} \\ \end{tabular} \normalsize The use of deictic verbs such as {\it kuru\/} (`come'), and {\it iku\/} (`go') also indicate speaker's perspective. Kuno calls a verb that is sensitive to the speaker's perspective an {\sc Empathy-loaded} verb, and defines {\sc Empathy locus} as the argument position whose referent the speaker is identifying with\footnote{The speaker does not necessarily take his/her own perspective to describe an event in which s/he is involved.}. Any Japanese verb can be made into an empathy-loaded verb by using an empathy-loaded verb as an auxiliary, which is suffixed onto the main verb stem. The complex predicate made by this operation inherits the empathy-locus of the suffixed verb. The {\em kureru} form of (`give') can be used as a suffix, to mark {\sc obj} or {\sc obj2} as the empathy-locus, as can the deictic verb {\em kuru\/} (`come') The use of the suffix {\em kureta} is shown in (3). \noindent (3)\\ \footnotesize \begin{tabular}{llllll} Hanako wa & Taroo ni & hon o & yonde-kureta. & & \\ & & book & read-gave & & \\ \multicolumn{6}{l}{{\it ``Hanako gave Taroo a favor of reading a book.''}}\\ \multicolumn{6}{l}{EMPATHY=OBJ2=TAROO} \\ \end{tabular} \normalsize The suffixation of verbs such as {\em iku\/} (`go') and the {\em yaru} form of (`give'), mark {\sc subject} as the empathy-locus, e.g. {\em itta} in (4). \noindent (4)\\ \footnotesize \begin{tabular}{llllll} Hanako wa & Taroo o & tazunete-itta. & & & \\ & & visit-went & & & \\ \multicolumn{6}{l}{{\it ``Hanako went to visit Taroo.''}}\\ \multicolumn{6}{l}{EMPATHY=SUBJ=HANAKO} \\ \end{tabular} \normalsize The relevance of speaker's empathy to centering is that a discourse entity realized as the empathy-locus is more salient, so that the empathy-locus position is ranked higher on the $\rm{Cf}$. Therefore, we use a ranking for the $\rm{Cf}$ in Japanese that incorporates {\sc Empathy} as follows: \begin{quote} {\bf Cf Ranking for Japanese} \\ {\sc topic} $>$ {\sc empathy} $>$ {\sc subj} $>$ {\sc obj2} $>$ {\sc obj} \end{quote} This ranking is a slight variation of that proposed by Kameyama\cite{Kameyama86a}. The centering algorithm works by taking the arguments of the verb and ordering them according to the Cf ranking for Japanese given above. In the cases where there are zero pronouns, there will be multiple possibilities for their interpretation and this will result in there being a priori several possible Cf lists\footnote{A discourse entity can simultaneously fulfill multiple roles. The entity is ranked according to the highest ranked role.}. These Cf lists are filtered according to the centering rules and constraints in section \ref{cent-form}. If there are still multiple possibilities, then the ordering on transitions applies, and {\sc continuing} interpretations are preferred. Many cases of the preference for one interpretation over another follow directly from the distinction between {\sc continuing} and {\sc retaining}. \noindent (5) \\ \footnotesize \begin{tabular}{llllll} $\rm{U_{n}}$: & & & & & \\ Taroo wa & paatii ni & syootai-sareta. & & &\\ & party to & invited-was & & &\\ \multicolumn{5}{l}{{\it ``Taroo was invited to the party.''}}&\\ \end{tabular} \begin{tabular}{|llllll|} \hline {\bf Cb:} & TAROO & & & & \\ {\bf Cf:} & [TAROO] & & & & \\ \hline \end{tabular} \begin{tabular}{llllll} $\rm{U_{n+1}}$: & & & & & \\ 0 & Hanako o & totemo & kiniitta. & & \\ & & very-much & was-fond-of & & \\ \multicolumn{5}{l}{{\it ``He liked Hanako very much.''}} &\\ \end{tabular} \begin{tabular}{|llllll|} \hline {\bf Cb:} & TAROO & & & & \\ {\bf Cf:} & [TAROO, & HANAKO] & & & \\ & subj & obj & & & \\ \hline \end{tabular} \begin{tabular}{llllll} $\rm{U_{n+2}}$: & & & & & \\ Kinoo & 0 & 0 & eiga ni & sasotta rasii.& \\ yesterday & & & movie to & invite seems & \\ \multicolumn{5}{l}{{\it ``Seemingly he invited her to a movie.''}} &\\ \end{tabular} \begin{tabular}{|llllll|} \hline {\bf Cb:} & TAROO & & & & \\ {\bf Cf1:} & [TAROO, & HANAKO] & CONTINUING & & \\ & subj & obj & & & \\ {\bf Cf2:} & [HANAKO, & TAROO] & RETAINING & & \\ & subj & obj & & & \\ \hline \end{tabular} \normalsize When the centering algorithm applies in (5) to $\rm{U_{n+2}}$, constraint 3 says the $\rm{Cb}$ must be the highest ranked element of Cf($\rm{U_{n+1}}$) realized in $\rm{U_{n+2}}$. Because there are 2 zeros in $\rm{U_{n+2}}$, TAROO must be realized and therefore must be the $\rm{Cb}$. The only {\sc continuing} interpretation available, {\em Taroo invited Hanako ...}, corresponds to the forward centers list Cf1. The fact that the preferred interpretation is the one in which the {\sc subject} zero pronoun takes a {\sc subject} antecedent is epiphenomenal. Example (6) demonstrates the effect of speaker's empathy on the salience of discourse entities. \noindent (6) \\ \nopagebreak \footnotesize \begin{tabular}{llllll} $\rm{U_{n}}$: & & & & & \\ Hanako wa & tosyokan de & benkyoositeita. & & &\\ & library in & studying-was & & &\\ \multicolumn{5}{l}{{\it ``Hanako was studying in the library.''}} &\\ \end{tabular} \begin{tabular}{|llllll|} \hline {\bf Cb:} & HANAKO & & & & \\ {\bf Cf:} & [HANAKO] & & & & \\ \hline \end{tabular} \footnotesize \begin{tabular}{llllll} $\rm{U_{n+1}}$: & & & & & \\ Taroo ga & Hanako o & tetudatte-kureta. & & & \\ & & help-gave & & & \\ \multicolumn{5}{l}{{\it ``Taroo gave Hanako a favor in helping her.''}} &\\ \end{tabular} \begin{tabular}{|llllll|} \hline {\bf Cb:} & [HANAKO] & & & & \\ {\bf Cf:} & [HANAKO, & TAROO] & & & \\ & empathy & subj & & & \\ \hline \end{tabular} \begin{tabular}{llllll} $\rm{U_{n+2}}$: & & & & & \\ Tugi no hi & 0 & 0 & eiga ni & sasotta. & \\ next of day &SUBJ & OBJ & movie to & invited & \\ \multicolumn{5}{l}{{\it ``Next day she invited him to a movie.''}}& \\ \end{tabular} \begin{tabular}{|llllll|} \hline {\bf Cb:} & HANAKO & & & & \\ {\bf Cf:} & [HANAKO, & TAROO] & CONTINUING & & \\ & subj & obj & & & \\ \hline \end{tabular} \normalsize In (6), HANAKO is the most highly ranked entity from $\rm{U_{n+1}}$ realized in $\rm{U_{n+2}}$ , and therefore must be the $\rm{Cb}$. The preferred interpretation will therefore be the {\em she invited him...} one that results from the more highly ranked {\sc continuing} transition, in which HANAKO is the preferred center ($\rm{Cp}$). The centering algorithm can also be applied successfully to intrasentential anaphora, by treating the subordinate clause as though it were a separate utterance for the purposes of pronoun interpretation. Consider: \noindent (7) \\ \footnotesize \begin{tabular}{llllll} Taroo wa & Kim ni &[0 0 & bengosuru] &koto o &hanasita. \\ & & & defend & comp & told \\ \multicolumn{6}{l}{{\it``Taroo told Kim that he would defend her''}}\\ \end{tabular} \begin{tabular}{|llllll|} \hline {\bf Cb:} & TAROO & & & & \\ {\bf Cf1:} & [TAROO, & KIM] & CONTINUING & & \\ & subj/top & obj2 & & & \\ {\bf Cf2:} & [HANAKO, & KIM] & RETAINING & & \\ & subj/top & obj2 & & & \\ \hline \end{tabular} \normalsize The {\sc continuing} interpretation, {\em Taroo told Kim that he would defend her}, is preferred to the {\sc retaining} interpretation, {\em Taroo told Kim that she would defend him}. \section{Topic ambiguity} The centering process reduces but does not necessarily eliminate semantic ambiguity in Japanese discourse. Within a loosely defined context, a native speaker's intuitions sometimes still allow for more than one equally preferred interpretation of an utterance. \subsection{Center Establishment} In the ``Introduce'' example shown in (8) below, ambiguity arises from the combined facts that the $\rm{Cb}$ of $\rm{U_{1}}$ is neutral (undefined), and there are more entities on the Cf list of $\rm{U_{1}}$ than there are zero pronouns in $\rm{U_{2}}$. \noindent (8)\\ \nopagebreak \footnotesize \begin{tabular}{llllll} $\rm{U_{1}}$: & & & & & \\ Lyn-ga & Masayo-ni & Sharon-o & shookaisita & & \\ SUBJ & OBJ2 & OBJ & introduced & & \\ \multicolumn{6}{l}{\it {``Lyn introduced Sharon to Masayo.''}}\\ \end{tabular} \begin{tabular}{|llllll|} \hline {\bf Cb:} & [?] & & & & \\ {\bf Cf:} & [LYN, & MASAYO, & SHARON] & & \\ & subj & obj2 & obj & & \\ \hline \end{tabular} \begin{tabular}{llllll} $\rm{U_{2}}$: & & & & & \\ 0 & 0 & kiniitteiru & & & \\ \multicolumn{6}{l}{\it {``Lyn likes Masayo''} (Cf1a)}\\ \multicolumn{6}{l}{\it {``Lyn likes Sharon''} (Cf1b)}\\ \multicolumn{6}{l}{\it {``Masayo likes Sharon''} (Cf2)}\\ \end{tabular} \begin{tabular}{|llllll|} \hline {\bf Cb1:} & LYN & & & & \\ {\bf Cb2:} & MASAYO & & & & \\ {\bf Cf1a:} & [LYN, & MASAYO] & & & \\ & subj & obj & & & \\ {\bf Cf1b:} & [LYN, & SHARON] & & & \\ & subj & obj & & &\\ {\bf Cf2:} & [MASAYO, & SHARON] & & & \\ & subj & obj & & &\\ \hline \end{tabular} \normalsize All three of these readings of $\rm{U_{2}}$ are equally preferred {\sc continuations}. To explain this fact, we posit that the $\rm{Cb}$ of an initial utterance $\rm{U_{n}}$ may be treated as a variable, indicated by [?], which can be equated with whatever $\rm{Cb}$ is assigned to the subsequent utterance $\rm{U_{n+1}}$\footnote{Future work will discuss center establishment in more detail, as well as other interactions, e.g., the effect of {\it wa} marking.}. For example, because there are 2 zeros in $\rm{U_{2}}$ of (8) and there are 3 entities available to fill these positions, constraint 3 implies that SHARON (the lowest ranked entity) can never be the $\rm{Cb}$, since it will never be the most highly ranked element of $\rm{Cf(U_1)}$ realized in $\rm{U_2}$. Therefore, whenever LYN is realized, the {\sc continuation} interpretation will place LYN in subject position, thus explaining the first two readings of $\rm{U_2}$. The third reading is available because no $\rm{Cb}$ has yet been established for $\rm{U_{1}}$, so that a {\sc continuation} does not require the realization of LYN in $\rm{U_{2}}$. Notice that any reading that assigns SHARON to the subject position or LYN to a non-subject position would produce a {\sc retention}. \subsection{Zero Topics} Another class of ambiguities can result from the optional assignment of {\sc topic} to a zero pronoun. We propose a topic assignment rule: \begin{quote} {\bf Zero Topic Assignment} \\ When no {\sc continuation} transition is available, and a zero pronoun in $\rm{U_{m}}$ represents an entity that was the $\rm{Cb(U_{m-1})}$ and if no other entity in $\rm{U_{m}}$ is overtly marked as the {\sc topic}, that zero may be interpreted as the {\sc topic} of $\rm{U_{m}}$. \end{quote} This fact, which has been overlooked in previous treatments of zero pronouns in Japanese, explains the interesting contrast between the two discourse segments in examples (9) and (10) below. Assume in (9) and (10) that TAROO and HANAKO have already been under discussion:\footnote{Due to lack of space, we can not discuss the interaction of center establishment with zero topic assignment here.} \noindent (9)\\ \footnotesize \begin{tabular}{llllll} $\rm{U_{n}}$: & & & & & \\ Taroo wa & kooen o & sanpo-siteita & & & \\ SUBJ & park & walk-around & & & \\ \multicolumn{6}{l}{\it {``Taroo was walking around the park''}}\\ \end{tabular} \begin{tabular}{|llllll|} \hline {\bf Cb:} & TAROO & & & & \\ {\bf Cf:} & [TAROO, & PARK] & & & \\ & subj & obj & & & \\ \hline \end{tabular} \begin{tabular}{llllll} $\rm{U_{n+1}}$: & & & & & \\ Hanako ga & 0 & yatto & mituketa & & \\ SUBJ & & finally & found & & \\ \multicolumn{6}{l}{\it {``Hanako finally found (him).''}}\\ \end{tabular} \begin{tabular}{|llllll|} \hline {\bf Cb:} & TAROO & & & & \\ {\bf Cf1:} & [TAROO, & HANAKO] & (C) & &\\ & topic/obj & subj & & & \\ {\bf Cf2:} & [HANAKO, & TAROO] & (R) & & \\ & subj & obj & & & \\ \hline \end{tabular} \begin{tabular}{llllll} $\rm{U_{n+2}}$: & & & & & \\ 0 & 0 & yotei o & setumeisita & & \\ SUBJ& OBJ & schedule & explained & & \\ \multicolumn{6}{l}{\it{He explained the schedule to her.} (Cf1)}\\ \multicolumn{6}{l}{\it{She explained the schedule to him.} (Cf2)}\ \end{tabular} \begin{tabular}{|llllll|} \hline {\bf Cb1:} & TAROO & & & & \\ {\bf Cb2:} & HANAKO & & & & \\ {\bf Cf1:} & [TAROO, & HANAKO] & (C) & & \\ & subj & obj & & &\\ {\bf Cf2:} & [HANAKO, & TAROO] & (S-1) & & \\ & subj & obj & & &\\ \hline \end{tabular} \normalsize In (9), there are actually two possible Cf lists in $\rm{U_{n+1}}$; Cf2, which is the only list possible without topic ambiguity, represents a {\sc retention} (R) rather than a {\sc continuation} (C), thus triggering zero topic assignment. The utterance $\rm{U_{n+1}}$, actually has the same meaning for both Cf lists. The ambiguity in $\rm{U_{n+2}}$ is caused by the fact that the hearer simultaneously entertains both of the $\rm{Cf(U_{n+1})}$. The availability of zero topic assignment means that TAROO can be the $\rm{Cp}$ even when TAROO is realized as the topic/object. The {\sc shift-1} interpretation results from the algorithm's application to Cf2 of $\rm{U_{n+1}}$. We can test to see if topic ambiguity is actually the discourse phenomenon at work here by contrasting (9) with its minimal pair (10), in which overt topic marking in $\rm{U_{n+1}}$ rules out topic ambiguity. \noindent (10)\\ \footnotesize \begin{tabular}{llllll} $\rm{U_{n}}$: & & & & & \\ Taroo wa & kooen o & sanpo-siteita & & & \\ SUBJ & park & walk-around & & & \\ \multicolumn{6}{l}{\it {``Taroo was walking around the park''}}\\ \end{tabular} \begin{tabular}{|llllll|} \hline {\bf Cb:} & TAROO & & & & \\ {\bf Cf:} & [TAROO, & PARK] & & & \\ & subj & obj & & & \\ \hline \end{tabular} \begin{tabular}{llllll} $\rm{U_{n+1}}$: & & & & & \\ Hanako-wa & 0 & yatto & mituketa & & \\ TOP/SUBJ & & finally & found & & \\ \multicolumn{6}{l}{\it {``Hanako finally found (him).''}}\\ \end{tabular} \begin{tabular}{|llllll|} \hline {\bf Cb:} & TAROO & & & & \\ {\bf Cf:} & [HANAKO, & TAROO] & (R) & & \\ & top/subj & obj & & & \\ \hline \end{tabular} \begin{tabular}{llllll} $\rm{U_{n+2}}$: & & & & & \\ 0 & 0 & yotei-o & setumeisita & & \\ & & schedule & explained & & \\ \multicolumn{6}{l}{\it {``She explained the schedule to him''}}\\ \end{tabular} \begin{tabular}{|llllll|} \hline {\bf Cb:} & HANAKO & & & & \\ {\bf Cf:} & [HANAKO, & TAROO] &(SHIFT1) & & \\ & subj & obj & & & \\ \hline \end{tabular} \normalsize In (10) the only Cf possible for $\rm{U_{n+1}}$ is the {\sc retention} in the parallel utterance in (9). Given that there are 2 zero pronouns in $\rm{U_{n+2}}$, constraint 3 forces a shift. The {\em Hanako explained ...} interpretation is preferred because it is the more highly ranked {\sc shift-1} transition. If {\sc hanako} could represent a {\sc topic-obj} there would be another equally ranked {\sc shift-1} interpretation. However, HANAKO can not be a zero topic because it was not the $\rm{Cb}$ of the previous utterance. \section{Discussion} We have demonstrated a computational treatment of the resolution of zero pronouns in Japanese. Kameyama proposed an analysis of Japanese zero pronouns that used centering, but did not distinguish between {\sc continuing} and {\sc retaining}, and thus required an extra mechanism, i.e. property-sharing\cite{Kameyama85}. Our examples (5), (6) and (7) show that property-sharing is an unnecessary stipulation. In addition, there are a number of cases in which property-sharing just doesn't work. Our ``introduce'' example (8) illustrates that it is not essential for a zero pronoun to share a grammatical function property with its antecedent. In fact property-sharing would falsely predict that the {\em Masayo likes Sharon} interpretation of (8) $\rm{U_2}$ is not possible, as well as falsely predicting the ungrammaticality of examples like (11) below. \noindent(11)\\ \footnotesize \begin{tabular}{llllll} $\rm{U_{n}}$: & & & & & \\ & Hanako wa & repooto o & kaita. & & \\ & & report & wrote & & \\ & \multicolumn{5}{l}{\it {``Hanako wrote a report''}}\\ & & & & & \\ $\rm{U_{n+1}}$: & & & & & \\ & $0_{i}$ & Taroo ni & aini-itta. & & \\ & & & to see-went & & \\ & \multicolumn{5}{l}{\it {``She went to see Taroo''}}\\ & \multicolumn{5}{l}{$0_{i}$ = Hanako [SUB EMPATHY]}\\ & & & & & \\ $\rm{U_{n+2}}$: & & & & & \\ & Taroo wa & $0_{i}$ & kibisiku hihansita. & & \\ & & & severely criticized & & \\ & \multicolumn{5}{l}{\it {``Taroo severely criticized her.''}}\\ & \multicolumn{5}{l}{$0_{i}$ = Hanako [nonSUB nonEMPATHY]}\\ \end{tabular} \normalsize Property-sharing requires that in $\rm{U_{n+2}}$, $i \neq $ HANAKO, since the zero carries the properties ({\sc subj}, {\sc empathy}) in $\rm{U_{n+1}}$, but has the properties ({\sc nonsubj},{\sc nonempathy}) in $\rm{U_{n+2}}$\footnote{Kameyama called the Empathy property IDENT.}. But in fact $\rm{U_{n+2}}$ is perfectly acceptable under the intended reading of {\em Taroo severely criticized Hanako}. Nothing special needs to be said about these to get the correct interpretation using the centering algorithm. We have also proposed a notion of topic ambiguity, which arises from the fact that the grammatical function of unexpressed zero arguments is indeterminate. The application of zero topic assignment also depends on the centering theory distinction between {\sc continuing} and {\sc retaining}. In addition, the centering construct of backward-looking center, $\rm{Cb}$, gives us a computational way of determining when a zero pronoun may be assigned {\sc Topic}. Topic ambiguity has been ignored in previous analyses, but it explains the availability of interpretations that previous accounts would predict as ungrammatical. This analysis has implications for the design of language-independent discourse processing modules. We claim that the syntactic factors that affect the ranking of the items on the forward center list, $\rm{Cf}$, will vary from language to language. The ordering for Japanese incorporates {\sc topic} and {\sc empathy} into the $\rm{Cf}$ ranking, which is a single parameter of the centering algorithm. In every other respect, the rules and constraints of the centering framework that the centering algorithm implements remain invariant. \section{Acknowledgements} The authors would like to thank Aravind Joshi, Carl Pollard, and Ellen Prince for their comments and support. This paper has also benefited from suggestions by Megan Moser, Peter Sells, Enric Vallduv\'{\i}, Bonnie Webber and Steve Whittaker. This research was partially funded by ARO grants DAAG29-84-K-0061 and DAAL03-89-C0031PRI, DARPA grant N00014-85-K0018, and NSF grant MCS-82-19196 at the University of Pennsylvania, and by Hewlett-Packard Laboratories.
proofpile-arXiv_065-457
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} The magnetic moments of the stable baryons are very well measured, and their theoretical calculation gives a sensitive test of our understanding of baryon structure in quantum chromodynamics (QCD). Although the basic pattern and approximate magnitudes of the moments can be explained using the nonrelativistic quark model, the deviation of the moments from the quark-model pattern has not been explained dynamically despite many attempts. We discuss that problem here from the point of view of nonperturbative QCD. In particular, we have analyzed the theory of the moments in the approximate QCD setting developed in the work of Brambilla {\em et al.} \cite{brambilla}, who used a Wilson-loop approach to study the $qqq$ bound state problem. By modifying their derivation of the $qqq$ interaction and wave equation, we have derived expressions for the baryon moments in terms of the underlying quark moments, including the first corrections associated with the binding of the quarks in baryons. Our results hold in the ``quenched'' approximation in which there are no internal quark loops embedded in the world sheet swept out by the Wilson lines joining the valence quarks, and no pairs associated with the valence lines. The corrections to the moments are of relative order $\langle V\rangle/m_q$, where $V$ is a typical component of the binding potential and $m_q$ is a constituent quark mass, and are potentially large enough to explain the deviations of the measured moments from the quark-model values. To test the theory, we have constructed variational wave functions for the baryons using the interactions derived by Brambilla {\em et al.} \cite{brambilla}, including all spin and orbital configurations possible for $J^P=\frac{1}{2}^+$ and internal and total orbital angular momenta $L\leq 2$, and used them to calculate the moments. The effects of excited orbital contributions to the moments are negligible, as expected. The new contributions tend to cancel, and the remnants do not have the correct pattern to explain the discrepancies between theory and experiment. The approximations underlying the Wilson-line analysis are known. Since of these, only the quenched approximation has a direct effect on the spin dependent terms with which we are dealing, we conclude that the use of the quenched approximation is responsible for the deviations of theory from experiment, and that the moment problem provides a sensitive test of this standard approximation in lattice and analytic QCD. In a world-sheet picture, the inclusion of internal quark loops which describe meson emission and absorption by the baryon would allow the introduction of internal orbital angular momentum and spin, and would affect the moments. We are now investigating meson-loop contributions using approximations suggested by the world-sheet picture and chiral pertubation theory. In the following sections, we sketch our derivations and calculations, and justify our conclusions. A more detailed discussion will be given elsewhere. \section{Baryon moments in QCD} \subsection{The problem} The simple, nonrelativistic quark model gives a qualitatively good description of the baryon moments. Under the assumption that each baryon is composed of three valence or constituent quarks in a state with all internal orbital angular momenta equal to zero, the moments are given by expectation values of the spin moment operators \begin{equation} \mbox{\boldmath$\mu$}_{\rm QM}=\sum_{\rm q}\,\mu_{\rm q} \mbox{\boldmath$\sigma$}_{\rm q}, \label{eq:mu} \end{equation} leading to the standard expressions \begin{equation} \mu_{\rm p}=\frac{1}{3}(4\mu_{\rm u}-\mu_{\rm d}),\,\ldots,\quad \mu_{\rm q}= \frac{e_{\rm q}}{2m_{\rm q}}\,. \label{eq:mu_quark} \end{equation} The masses in the quark moments are clearly effective masses, and can be treated as free parameters in attempting to fit the data. A least-squares fit to the measured octet moments, taken with equal weights, is shown in Table 1. The moments are given in nuclear magnetons (nm). The pattern of the signs of the quark model moments agrees with observation, while the root-mean-square deviation of theory from experiment is 0.14 nuclear magnetons, or about 9\% of the average magnitudes of the moments. Agreement at this level can be regarded as an outstanding success of the quark model, but the deviations also give a very sensitive test of baryon structure: there is presently no completely successful first-principles theory of the moments. \bigskip \begin{center} Table 1: Quark model fit to the magnetic moments in the baryon octet. All moments are given in nuclear magnetons. \medskip \begin{tabular}{|c|c|c|r|} \hline Baryon & Experiment & Quark Model & $\Delta\mu$\\ \hline $p$ & 2.793 & 2.728 & 0.065 \\ $n$ & -1.913 & -1.819 & -0.094 \\ $\Sigma^+$ & $2.458\pm 0.010$ & 2.639 & -0.181 \\ $\Sigma^-$ & $-1.160\pm 0.025$ & -0.999 & -0.161 \\ $|\Sigma^0\rightarrow\Lambda\gamma|$ & $1.61\pm 0.08$ & 1.575 & -0.03 \\ $\Lambda$ & $-0.613\pm 0.004$ & -0.642 & 0.029 \\ $\Xi^0$ & $-1.250\pm 0.014$ & -1.462 & 0.212 \\ $\Xi^-$ & $-0.651\pm 0.025$ & -0.553 & -0.098 \\ \hline \end{tabular} \end{center} \medskip \subsection{The Wilson-loop approach to the baryon moments} Our approach to the baryon moment problem is based on the work of Brambilla {\em et al.} \cite{brambilla}, who derived the interaction potential and wave equation for the valence quarks in a baryon from QCD using a Wilson-line construction. The basic idea is to construct a Green's function for the propagation of a gauge-invariant combination of quarks joined by path ordered Wilson-line factors \begin{equation} U=P\exp ig\!\int\! A_g\cdot dx, \label{wilsonline} \end{equation} where $A_g$ is the color gauge field. The Wilson lines sweep out a three-sheeted world sheet of the form shown in Fig.\ \ref{fig:worldsheet} as the quarks move from their initial to their final configurations. By making an expansion in powers of $1/m_q$ and considering only forward propagation of the quarks in time using the Foldy-Wouthuysen approximation, Brambilla {\em et al.} are able to derive a Hamiltonian and Schr\"{o}dinger equation for the quarks, with an interaction which involves an average over the gauge field. That average is performed using the minimal surface approximation in which fluctuations in the world sheet are ignored, and the geometry is chosen to minimize the total area of the world sheet {\em per} Wilson, subject to the motion of the quarks. The short-distance QCD interactions are taken into account explicitly. \begin{center} \begin{figure}[t] \psfig{figure=worldsheetfig.eps,height=1.9in} \caption{World sheet picture for the structure of a baryon. \label{fig:worldsheet}} \end{figure} \end{center} The result of this construction is an effective Hamiltonian to be used in a semirelativistic Schr\"{o}dinger equation $H\Psi=e\Psi$, \begin{eqnarray} H &=& \sum_i\sqrt{p_i^2+m_i^2}+\sigma(r_1+r_2+r_3)-\sum_{i<j}\frac{2}{3} \frac{\alpha_{\rm s}}{r_{ij}} \nonumber \\ &&-\frac{1}{2m_1^2}\frac{\sigma}{r_1}{\bf s}_1 \cdot ({\bf r}_1\times{\bf p}_1) +\frac{1}{3m_1^2}{\bf s}_1\cdot[({\bf r}_{12}\times{\bf p}_1) \frac{\alpha_{\rm s}}{r_{12}^3} + ({\bf r}_{13}\times{\bf p}_1) \frac{\alpha_{\rm s}}{r_{13}^3}] \nonumber\\ &&-\frac{2}{3}\frac{1}{m_1m_2}\frac{\alpha_{\rm s}}{r_{12}^3} {\bf s}_1\cdot{\bf r_{12}}\times{\bf p}_2 -\frac{2}{3}\frac{1}{m_1m_3}\frac{\alpha_{\rm s}}{r_{13}^3} {\bf s}_1\cdot{\bf r_{13}}\times{\bf p}_3+\cdots. \label{hamiltonian} \end{eqnarray} Here ${\bf r}_{ij}={\bf x}_i-{\bf x}_j$ is the separation of quarks $i$ and $j$, $r_i$ is the distance of quark $i$ from point at which the sum $r_1+r_2+r_3$ is minimized, and ${\bf p}_i$ and ${\bf s}_i$ are the momentum and spin operators for quark $i$. The parameter $\sigma$ is a ``string tension'' which specifies the strength of the long range confining interaction, and $\alpha_{\rm s}$ is the strong coupling. The terms hidden in the ellipsis include tensor and spin-spin interactions which will not play a role in the analysis of corrections to the moment operator, and the terms that result from permutations of the particle labels. The full Hamiltonian is given in \cite{brambilla}. This Hamiltonian, including the terms omitted here, gives a good description of the baryon spectrum as shown by Carlson, Kogut, and Pandharipande \cite{kogut} and by Capstick and Isgur \cite{isgur}, who proposed it on the basis of reasonable physical arguments, but did not give formal derivations from QCD. The presence of the quark momenta ${\bf p}_i$ in the Thomas-type spin-dependent interactions in Eq.\ (\ref{hamiltonian}) suggests that new contributions to the magnetic moment operator could arise in a complete theory through the minimal substitution \begin{equation} {\bf p}_i\rightarrow {\bf p}_i-e_i{\bf A}_{\rm em}(x_i), \label{minimalsub} \end{equation} with ${\bf A}_{\rm em}(x_i)$ the electromagnetic vector potential associated with an external magnetic field. However, this point is obscured in the work of Brambilla {\em at al.} by the transformations they make to express the equations for the Green's function in Wilson-loop form. We have therefore repeated their derivation of the valence-quark Hamiltonian in Eq.\ ( \ref{hamiltonian}), replacing the SU(3)$_c$ color gauge field $gA_g$ by the full SU(3)$_c\times$U(1)$_{\em em}$ gauge field $gA_g+e_qA_{\rm em}$ at the beginning. By then reorganizing the calculation of the three-quark Hamiltonian to keep $e_qA_{\rm em}$ explicit throughout, and then expanding to first order in $A_{\rm em}$ with \begin{equation} {\bf A}_{\rm em}=\frac{1}{2}{\bf B}\times{\bf x}_q, \quad {\bf B}={\rm constant,} \label{magneticpotential} \end{equation} we can identify the modified magnetic moment operator through the relation \begin{equation} \Delta H=-\mbox{\boldmath$\mu$}\cdot{\bf B}. \label{deltaH} \end{equation} The new moment operator, $\mbox{\boldmath$\mu$}=\mbox{\boldmath$\mu$}_{\rm QM}+\Delta\mbox{\boldmath$\mu$}$, involves the leading corrections to the quark-model operator associated with the binding interactions. $\Delta\mbox{\boldmath$\mu$}$ {\em can}, in fact, be read off from the terms in Eq.\ (\ref{hamiltonian}) which depend on both the quark spins and momenta by making the minimal substitution in Eq.\ (\ref{minimalsub}). For example, the term which depends on ${\bf s}_1\cdot{\bf r}_{12}\times {\bf p}_1$ gives an extra contribution \begin{equation} \frac{e_1}{6m_1^2}{\bf x}_1\times({\bf s}_1\times{\bf r}_{12})\frac{\alpha_ {\rm s}}{r_{12}^3} \label{deltamu1} \end{equation} to $\mbox{\boldmath$\mu$}_1$. There are also possible orbital contributions to the moments because the Hamiltonian mixes states with nonzero orbital angular momenta with the ground state. These have the standard form to the accuracy we need. The baryon moments are now given by expressions of the form \begin{equation} \mu=\sum_{i=1}^3\mu_i\langle\sigma_{i,z}\rangle(1+\delta_i)+\sum_{i=1}^3 \mu_i\langle L_{i,z}\rangle, \label{mufinal} \end{equation} where we have quantized along $\bf B$, taken along the $z$ axis. The expectation values are to be calculated in the baryon ground states. The correction terms $\delta_i$ from the new operators are given for the $L=0$ baryons other than the $\Lambda$ by \begin{eqnarray} \delta_i &=& \frac{3\epsilon_0+\epsilon_1}{2m_1}+\frac{e_3}{e_1} \frac{\epsilon_2}{m_3}-\frac{\Delta_0+\Delta_1}{2m_1},\quad i=1,2,\nonumber\\ \delta_3 &=& \frac{\epsilon_2}{m_3}+\frac{2e_1}{e_3}\frac{\epsilon_1}{m_1} -\frac{\Delta_2}{m_3},\label{deltas} \end{eqnarray} where the $\epsilon$'s and $\Delta$'s are ground state radial matrix elements, \begin{eqnarray} \epsilon_0=\langle\frac{2}{3}\frac{\alpha_{\rm s}}{6r_{12}}\rangle,\quad & \displaystyle \epsilon_1=\langle\frac{2}{3}\frac{\alpha_{\rm s} {\bf r}_{23}\cdot{\bf z}_2}{3r_{23}^3}\rangle,\quad & \epsilon_2= \langle\frac{2}{3}\frac{\alpha_{\rm s}{\bf r}_{31}\cdot {\bf z}_3}{3r_{31}^3}\rangle \nonumber\\ \Delta_0=\langle\frac{\sigma r_{12}}{12}\rangle,\quad & \displaystyle \Delta_1=\langle\frac{\sigma {\bf r}_{23}\cdot{\bf z}_2}{6r_{23}}\rangle,\quad & \Delta_2=\langle\frac{\sigma {\bf r}_{31}\cdot{\bf z}_3}{6r_{31}}\rangle. \label{matrixelements} \end{eqnarray} The identical quarks in these baryons are labelled 1 and 2, the unlike quark, 3. In writing these results, we have made the approximation $r_1+r_2+r_3\approx \frac{1}{2}(r_{12}+r_{23}+r_{31})$, known to be reasonably accurate for the ground state baryons \cite{kogut}, and used the corresponding Thomas spin interaction. The result for the $\Lambda$ is similar. \subsection{Tests of the model} Rough estimates of the matrix elements above suggest that $\epsilon_i/m_l \approx 0.05$ and that $\Delta_i/m_l\approx 0.3$ for the light-quark masses used in Refs.\ 2 and 3. Binding effects are therefore potentially large, and are different for different quarks and baryons. To obtain a quantitative test of these effects and reliable estimates of the orbital contributions to the moments, expected to be small, we have performed a detailed analysis of the baryon wavefunctions using the Hamiltonian given in \cite{brambilla}. We use Jacobi-type internal coordinates \mbox{\boldmath $\rho, \lambda$}, with $\mbox{\boldmath $\rho$}= {\bf x}_1-{\bf x}_2$, and $\mbox{\boldmath $\lambda$}={\bf R}_{12}-{\bf x}_3$, where ${\bf R}_{ij}$ is the coordinate of the center of mass of quarks $i$ and $j$. The most general $j=\frac{1}{2}^+$ baryon wave function for quarks 1 and 2 identical and orbital angular momenta $L_\rho, L_\lambda, L \leq 2$ has the form \begin{eqnarray} \psi_{\frac{1}{2},m}&=&\left[\right.a_0\psi_a{\bf 1} +ib_0\psi_b\,(\mbox{\boldmath$\sigma$}_1-\mbox{\boldmath$\sigma$}_2) \cdot\mbox{\boldmath$\rho$}\times\mbox{\boldmath$\lambda$} +c_0\psi_c\,t_{12}(\mbox{\boldmath$\rho$})\nonumber\\ &&+d_0\psi_d\,t_{12}(\mbox{\boldmath$\lambda$}) +e_0\psi_d\,(\mbox{\boldmath$\sigma$}_1-\mbox{\boldmath$\sigma$}_2)\cdot \mbox{\boldmath$\sigma$}_3\,\mbox{\boldmath$\rho$}\cdot \mbox{\boldmath$\lambda$}\left.\right]\chi_{\raisebox{-.5ex} {$\scriptstyle\frac{1}{2},m$}}^{S_{12}=1}. \label{wavefunc} \end{eqnarray} Here $t_{12}$ is the usual tensor operator \begin{equation} t_{12}({\bf x})=3\mbox{\boldmath$\sigma$}_1\cdot{\bf x}\mbox{\boldmath$\sigma$} _2\cdot{\bf x}-\mbox{\boldmath$\sigma$}_1\cdot\mbox{\boldmath$\sigma$}_2 {\bf x}^2, \end{equation} and $\chi_{\raisebox{-.5ex} {$\scriptstyle\frac{1}{2},m$}}^{S_{12}=1}$ is the standard three-particle spinor for $S_{12}=1,\,j=\frac{1}{2},\, j_3=m$. The scalar functions $\psi_i=\psi_i(\rho^2,\lambda^2)$ are normalized together with the accompanying spin operators. The constant coefficients $a_0,\ldots,e_0$, normalized to unity, give the fractions of the various component states in $\psi$. With this form of the wave function, we can use trace methods for the spins to calculate such quantities as $\langle H\rangle$ and $\langle\mbox{\boldmath$\mu$}\rangle$. We have made variational calculations of the energies and approximate wave functions of the ground state baryons and their first excited states using the Hamiltonian in \cite{brambilla}. The information on the excited states allows us to calculate the coefficients $b_0,\ldots, e_0$ perturbatively. While the Hamiltonian does mix orbitally excited states into the $L_\rho=L_\lambda=L=0$ quark-model ground state, the coefficients are very small, ranging from essentially zero to about 0.02 depending on the baryon. Because these coefficients only appear quadratically in the baryon moments, the orbital contributions to the moments are completely negligible. The radial matrix elements $\epsilon_i$ and $\Delta_i$ defined in Eq.\ (\ref{matrixelements}) are significant, with the $\epsilon$'s ranging from 10 to 20 MeV, and the $\Delta$'s from 40 to 70 MeV. A new fit to the moment data using the expressions in Eqs.\ (\ref{mufinal}) and (\ref{deltas}), with the quark masses allowed to vary, gives a small improvement in the fitted moments, with the root-mean-square deviation from the measured moments decreasing from 0.14\,nm for the quark model to 0.10\,nm. A secondary effect of the corrections is to change the fitted quark masses or moments significantly, with the effective quark masses decreasing, or the moments increasing. We have concluded from this exercise first, that the finer details of baryon structure are not yet described correctly in the present QCD-based model, and second, that the baryon magnetic moments provide a very sensitive check on the theory. \section{Possible improvements in the theory} We begin the discussion of possible improvements in the theory of the moments by recalling the approximations used in the construction given by Brambilla {\em et al.} \cite {brambilla} and in our derivation of the moment operators: \newcounter{approx} \begin{list}{(\roman{approx})}{\usecounter{approx}} \item the $1/m_q$ expansion; \item the minimal surface approximation; \item forward propagation of the quarks in time; \item and the quenched approximation. \end{list} Of these approximations, only (iii) and (iv) are likely to affect the moments significantly. The $1/m_q$ expansion is to be interpreted as an expansion in constituent quark masses. It can be resummed in the kinetic terms in the Hamiltonian as in Eq.\ (\ref{hamiltonian}), and leads elsewhere to the appearance of effective inverse masses $1/m_i$ which need not be the same as the kinematic masses, but represent averages of quantities such as $1/E_i=1/\sqrt{p_i^2+m_i^2}$. Since there is no new spin dependence involved, and we have treated the quark masses as free parameters in fitting the moments, relativistic corrections of this sort are unlikely to change our results. The average over the color gauge field $A_g$ in the minimal surface approximation neglects fluctuations, and minimizes the surface energy of the world sheet to obtain the approximate Hamiltonian. Since the color fields do not carry charge, they do not contribute internal currents in the baryon, and an improved treatment of the averaging would presumably not change the moment operator directly. While it could change the functional form of the Hamiltonian somewhat, the quantitative success of the model in describing baryon spectra suggests that the changes would not be large, and their effect on the moments through the $\epsilon$ and $\Delta$ parameters would be minimal. In contrast, the remaining two approximations affect the moment operator directly. Internal quark loops can contribute circulating currents in the baryon. These are omitted in the quenched approximation. In addition, the forward propagation of the quarks in time inherent in the use of the Foldy-Wouthuysen approximation eliminates quark pair effects connected with the valence lines in Fig. 1. We believe that this general quenched picture underlies the difficulties with the model, and that it will be necessary to include pair effects to obtain a precise dynamical description of the moments. One effect of internal quark loops can be seen in Fig.\ \ref{fig:mesonloop}. In this figure, we suppose that a quark loop is embedded in the minimal world sheet of the baryon. The effect is to give a meson state and a new baryon in a world-sheet picture of meson emission and absorption. \begin{figure}[h] \psfig{figure=mesonloopfig.eps,height=1.3in} \caption{Quark loops embedded in the world sheet give meson-baryon states. \label{fig:mesonloop}} \end{figure} \noindent An external magnetic field interacting with the system will see a meson current as well as the baryon current, and the moment will be modified. Meson currents were invoked in the past in attempts to explain the {\em full}\/ anomalous moments of the nucleons. Here we are only concerned with presumably small corrections to the quark-model moments. One approach to the calculation of meson loop effects is through chiral perturbation theory. This has a long history, and has been studied recently in the context of baryon moments by a number of authors \cite{jenkins,luty,bos}. The relevant diagrams for the interaction of the baryons with the pseudo Goldstone bosons of the chiral theory are shown in Fig.\ \ref{fig:goldstone}. \begin{figure}[h] \psfig{figure=goldstonefig.eps,height=1.2in} \caption{Goldstone-boson diagrams which contribution to the magnetic moments. \label{fig:goldstone}} \end{figure} The work of Jenkins {\em et al.} \cite{jenkins} uses heavy baryon chiral perturbation theory \cite{manohar}, and studies the effect on the moments of the ``nonanalytic terms'' in the symmetry breaking parameter $m_{\rm s}$. The analytic terms in $m_{\rm s}$ are ambiguous because the inevitable appearance of new couplings at each order in the chiral expansion, and are ignored. The results of Jenkins {\em et al.} are not encouraging as they stand, but we have found errors in some of the coupling factors given in the published paper. Luty {\em et al.} \cite{luty} concentrate on a simultaneous expansion in $1/N_{\rm c}$ and $m_{\rm s}$, and obtain interesting sum rules for the moments but little dynamical information. Finally, Bos {\em et al.} \cite{bos} consider the moments from the point of view of flavor SU(3) breaking in the baryon octet using a different chiral counting than that usually used, and obtain a very successful parametrization for the moments. However, this model is again nondynamical, and has seven parameters to describe eight measured moments. In fact, a fundamental problem with the chiral expansion from our perspective is that it simply parametrizes the moments with an expansion consistent with QCD, but does not control the higher order terms in what appears to be a slowly convergent series. We are presently investigating meson loop effects using a somewhat different approach suggested by the world-sheet picture. The baryon appears in this picture as an extended object which must absorb the recoil momentum in the emission of a meson. We would therefore expect wave function effects (form factors) to be important for high meson momenta, and to supply a natural cutoff for loop graphs. The wave function appears naturally when the process is viewed using ``old fashioned'' instead of Feynman perturbation theory, as indicated in Fig.\ \ref{fig:wavefunctionvertex}: energy \begin{figure} \psfig{figure=wavefunctionvertex.eps,height=1.4in} \caption{(a) Appearance of Bethe-Salpeter wave functions in the diagrams. (b) Schematic reminder of the distributed nature of the meson-baryon vertex. \label{fig:wavefunctionvertex}} \end{figure} denominators and vertex functions combine to give the $B'M$ component of the wave function for a baryon $B$, or, with different time ordering, the $BM$ component of the wave function of $B'$. Since the wave functions are expected to be fairly soft, with characteristic momenta below the chiral cutoff of $\sim 1$ GeV, they can be expected to combine a number of terms in the chiral expansion in a way that is dynamically accessible through approximate models derived from QCD such as that of Brambilla {\em et al.} \cite{brambilla}. The non-point character of the meson-baryon vertex in spacetime indicated in Fig.\ \ref{fig:wavefunctionvertex} contrasts sharply with the point vertex used in chiral perturbation theory, Fig.\ \ref{fig:goldstone}, and is also likely to play a role. We have reached the same conclusions by studying the exact sideways dispersion relations for the moments given by Bincer \cite{bincer}. The problem there is in extracting the quark-model moments. We note finally that a different aspect of symmetry breaking, the suppression of strange-quark loops through mass effects, appears as a natural possibility in a world-sheet picture. Note that this is {\em not} the same as the suppression of kaon loops considered by other authors; see, e.g., \cite{jenkins} and the references given there. \section{Conclusions} On the basis of the work sketched above, we have concluded that the magnetic moments of the baryons give a sensitive test of baryon structure and of approximations in QCD. In particular, the quenched approximation, while reasonably successful when used in the calculation of baryon and meson spectra, appears to fail for the moments. The accurate calculation of the moments by lattice methods will presumably involve going beyond that approximation, and will provide a precision test of the methods used. We find also that the world-sheet picture of baryon structure gives useful insights into the moments problem, and provides a new point of view which could be developed further. Problems which need further study within this approach to QCD include the following: \begin{list}{(\roman{approx})}{\usecounter{approx}} \item establishing the connection to, and the relevance of, the chiral limit; \item the development of methods to incorporate loop effects which build in the extended structure of the baryons; \item and the possible usefulness of string theory methods in the calculation of loop effects. \end{list} \section*{Acknowledgments} The authors have benefited from conversations with Dr.\ Nora Brambilla, and appreciate her interest in this work, and her organization of the Como Conference. This research was supported in part by the U.S. Department of Energy under Grant No.\ DE-FG02-95ER40896, and in part by the University of Wisconsin Graduate School with funds from the Wisconsin Alumni Research Foundation. \section*{References}
proofpile-arXiv_065-458
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} There is a wide class of integrable models which describe the flows between two different models of Conformal Field Theory (CFT) in UV and IR regions \cite{zam}. To describe these models the factorized scattering of massless particles was proposed in \cite{zz1}. In spite of general difficulties arising from the fact that the scattering of massless particles is not properly defined in 2D the application of this method provided very good results for several particular models. Certainly, the physical results are related to possibility of extracting the off-shell information from these S-matrices. Similarly to the massive case there are two ways to do that. The first way is TBA approach. The TBA equations can be written for the massless scattering which allow to calculate the effective central charge and to show that the latter really interpolates between corresponding UV and IR values \cite{zz1}. The second way consists in generalization proposed in \cite{muss} of the form factor bootstrap approach (which is originally formulated for the massive models \cite{book}) to the massless flows. One must distinguish between the form factor bootstrap in massive and massless cases. In the massive case the form factor bootstrap stays on the solid ground because the space of states is well defined as the Fock space of particles. In the massless case this very definition is doubtful and one has to consider the form factor bootstrap rather as intuitive than rigorous procedure. Indeed for certain operators (as, for example, the order-disorder operators for the flow from tricritical to critical Ising model) the straightforward application of the method leads to divergent series for the Green functions \cite{muss}. However, the situation can be improved even for those operators, furthermore there are operators for which the series converge providing spectacular examples of interpolation between UV and IR limits \cite{muss}. So, the form factor bootstrap is a method which works in massless case in spite of all the problems of general character. It must be noticed, however, that the complete construction of all the form factors is not given in the paper \cite{muss} even for the simplest model describing the flow from tricritical to critical Ising model considered there. In the present paper we shall give the complete construction for more complicated model: the Principal Chiral Field with Wess-Zumino-Novikov-Witten term. We shall explain briefly that our construction can be generalized to a wide class of models including the one considered in \cite{muss}. \section{Formulation of the problem.} On of the most beautiful examples of the massless flows for which the S-matrix is known is given by the Principal Chiral Field with Wess-Zumino-Novikov-Witten (WZNW) term on level $1$ (PCM${}_1$): \begin{eqnarray} S={1\over 2\lambda ^2}\int tr(g^{-1}\partial_{\mu}g) (g^{-1}\partial_{\mu}g) d^2x\ +\ i\Gamma (g) \nonumber \end{eqnarray} where the WZNW term $\Gamma (g)$ is defined by means of continuation of $g$ to 3D manifold $B$ for which the 2D space-time is the boundary: \begin{eqnarray} \Gamma (g)=\int\limits _{B} \epsilon ^{\mu \nu \lambda} tr(g^{-1}\partial_{\mu}g) (g^{-1}\partial_{\nu}g)(g^{-1}\partial_{\lambda}g)d^3x \nonumber \end{eqnarray} In the UV limit the pure PCM action dominates the central charge being equal to 3. In the IR region the flow is attracted by the fixed point $\lambda ^2=8\pi$ which corresponds to the conformal WZNW model with the central charge equal to 1. It is explained in \cite{zz1} that the flow arrives at the IR point along the direction defined by the irrelevant operator $\overline{T}T$ composed of the right and left components of the energy-momentum tensor. In the IR region the theory is conformal, so, two charialities essentially decouples. One describes corresponding left and right level-1 WZNW models in terms of massless particles. This is exactly left-left and right-right scattering which seems to be doubtful in 2D. The prescription of the paper \cite{zz1} for the definition of this scattering can be understood in the following way. We know that the level-1 WZNW model coincides with the UV limit of the massive $SU(2)$-invariant Thirring model. The local operators for the latter model are defined via their form factors \cite{book}. For the operators chiral in the limit (as chiral components of the energy-momentum tensor) the limiting values of the correlation functions are obtained by replacing the massive dispersion by the massless ones. On the other hand these limiting correlation functions coincide with the conformal ones. So, we get the representation of the conformal correlation functions for chiral operators in terms of the form factor series with massless particles. The form factors are defined through the S-matrix for the original massive theory which is considered now as S-matrix of massless particles. It is proposed to use this representations of the correlators as a starting point of the description of massless flows. More precisely, left and right particles are parametrized by the rapidities $\beta$ and $\beta'$ such that the energy-momentum are respectively $$e=-p=Me^{-\beta},\qquad e=p=Me^{\beta'} $$ where $M$ is the mass scale which can be chosen arbitrary on this stage. The theory possesses $SU(2)_L\otimes SU(2)_R$ symmetry, the left (right) movers are doublets with respect to $SU(2)_L$ ($SU(2)_R$). The factorizable S-matrices which describe the left and right CFT are given by \begin{eqnarray} S_{LL}(\beta _1,\beta_2)=S^Y(\beta _1-\beta_2),\qquad S_{RR}(\beta _1',\beta_2')=S^Y(\beta _1'-\beta_2') \nonumber \end{eqnarray} where $S^Y(\beta)$ is the Yangian S-matrix for the scattering of spin-${1\over 2}$ particles \cite{zz}: \begin{eqnarray} S^Y(\beta)={\Gamma ({1\over2}+{\beta\over 2\pi i})\Gamma (-{\beta\over 2\pi i})\over \Gamma ({1\over2}-{\beta\over 2\pi i})\Gamma ({\beta\over 2\pi i}) } \({\beta I-\pi i P\over \beta-\pi i}\) \nonumber \end{eqnarray} where $I$ and $P$ are respectively unit and permutation operators acting in the tensor product of two 2-dimensional isotopic spaces. The crucial point is in introducing the non-trivial left-right and right-left S-matrices. Contrary to $S_{LL}$ and $S_{RR}$ whose definition is rather formal the S-matrices $S_{LR}$ and $S_{RL}$ allow quite rigorous interpretation. For the PCM${}_1$ the proposal of \cite{zz1} is $$S_{RL}(\beta'-\beta)={1\over S_{LR}(\beta-\beta')}=U(\beta'-\beta),\qquad U(\beta)=\tanh{1\over2}\(\beta -{\pi i\over 2}\) $$ The scale normalization $M$ is fixed by the requirement that the zero of this S-matrix is situated exactly at $\beta ={\pi i\over 2}$. It is quite amusing that the IR limit corresponds to $\beta-\beta '=\log \Lambda$, $\Lambda\to\infty$, indeed in this limit the $s$-variable goes to zero. This fact is very interesting because in the massive case infinite rapidities are always related to UV behaviour of the form factors which has been investigated in several cases in \cite{book}, so, we can use the familiar methods for solving absolutely different problems. Let us describe the form factor bootstrap approach to massless flows as it is formulated in \cite{muss}. Consider the matrix element of certain local operator $\CO$ taken between the vacuum and the state containing The left and right particles with rapidities $\beta_1,\cdots,\beta_l$ and $\beta_1',\cdots,\beta_k'$ respectively: \begin{eqnarray} f_{\CO}(\beta_1,\cdots,\beta_l\ |\ \beta_1',\cdots,\beta_k') \label{Fb} \end{eqnarray} It is very convenient to collect all the rapidities together into the set $\theta _1,\cdots,\theta _{k+l}=$ $\beta_1,\cdots,\beta_l,\beta_1',\cdots,\beta_k'$ and to introduce index $a_i=L,R$ which distinguish the left and right particles. The first requirement of the form factors is that of symmetry: \begin{eqnarray} f_{\CO}(\cdots ,\theta _i,\theta _{i+1},\cdots)_{\cdots,a_i,a_{i+1},\cdots} S(\theta _i-\theta _{i+1})_{a_i,a_{i+1}} = f_{\CO}(\cdots ,\theta _{i+1},\theta _i,\cdots)_{\cdots,a_{i+1},a_i,\cdots} \label{a1} \end{eqnarray} If $a_i\ne a_j$ this equation has to be considered as definition which allows to construct the form factor with arbitrary placed left and right particles starting from (\ref{Fb}). The second requirement is \begin{eqnarray} &&f_{\CO}(\theta _1,\cdots ,\theta _{k+l-1},\theta _{k+l}+2\pi i) _{a_1,\cdots ,a_{k+l-1},a_{k+l}} = f_{\CO}(\theta _{k+l},\theta _1,\cdots ,\theta _{k+l-1}) _{a_{k+l},a_1,\cdots ,a_{k+l-1}}= \nonumber \\&&= f_{\CO}(\theta _1,\cdots ,\theta _{k+l-1},\theta _{k+l}) _{a_1,\cdots ,a_{k+l-1},a_{k+l}} S_{a_{k+l-1},a_{k+l}}(\theta _{k+l-1}-\theta _{k+l}) \cdots S_{a_{1},a_{k+l}}(\theta _{1}-\theta _{k+l}) \label{a2} \end{eqnarray} Since we do not have bound states in the theory the form factor $f_{\CO}(\theta _1,\cdots ,\theta _{k+l})$ is supposed to be a meromorphic function of $\theta _{k+l} $ in the strip $0<\theta _{k+l} <2\pi$ whose only singularities are the simple poles at the points $\theta _{k+l} =\theta _{j} +\pi i$. It is important that these singularities appear only in left-left and right-right chanels. The residue at $\theta _{k+l} =\theta _{k+l-1} +\pi i $ is given by \begin{eqnarray} &&2\pi i\ res f_{\CO}(\theta _1,\cdots ,\theta _{k+l-2},\theta _{k+l-1},\theta _{k+l}) _{a_1,\cdots ,a_{k+l-2} ,a_{k+l-1},a_{k+l}}= \nonumber \\&&\hskip 1cm = \delta _{a_{k+l-1},a_{k+l}}f_{\CO}(\theta _1,\cdots ,\theta _{k+l-2}) _{a_1,\cdots ,a_{k+l-2}}\otimes c_{k+l-1,k+l} \nonumber \\&&\hskip 1cm \times \(1\ -\ S_{a_{k+l-1},a_{1}}(\theta _{k+l-1}-\theta _{1})\cdots S_{a_{k+l-1},a_{k+l-2}}(\theta _{k+l-1}-\theta _{k+l-2})\) \label{a3} \end{eqnarray} here $c_{k+l-1,k+l}$ is a vector in the tensor product of two isotopic spaces constructed from the charge conjugation matrix, in our case it is the singlet vector in the tensor product of two spin-${1\over 2}$ representations of $SU(2)$. These requirements on the massless form factors do not differ too much from the form factor axioms of \cite{book}. However, the physical situation is quite different and the solutions to these equation can not be found in \cite{book}. \section{Form factors of the energy-momentum tensor.} It the present paper we are going to construct the form factors of the trace of energy-momentum tensor ($\Theta$) for PCM${}_1$. Our methods are applicable to other operators, but we are considering this particular one because of its nice properties and physical importance. Since the symmetry under $SU(2)_L\otimes SU(2)_R$ is not broken by the perturbation the form factors have to be singlets with respect to both isotopic groups. That is why $l=2n$ and $k=2m$. The form factors of $\Theta$ satisfy general conditions (\ref{a1}, \ref{a2}, \ref{a3}) and additional requirements following from the fact that we consider this particular operator. 1. The energy-momentum conservation implies that \begin{eqnarray} f_{\Theta}(\beta_1,\cdots,\beta_{2n}\ |\ \beta_1',\cdots,\beta_{2m}')= (\sum e^{-\beta _j})(\sum e^{\beta _j '}) f(\beta_1,\cdots,\beta_{2n}\ |\ \beta_1',\cdots,\beta_{2m}') \label{c1} \end{eqnarray} where for $n>1$ and $m>1$ the function $f$ does not other singularities that those of $f_{\Theta}$, for $n=1$($m=1$) it has simple poles at $\beta _2=\beta _1 +\pi i$ ($\beta _2 '=\beta _1 '+\pi i$) which are cancelled by $e^{-\beta _1 } +e^{-\beta _2 }$ ($e^{\beta _1' } +e^{\beta _2 '}$ ). 2. The lowest form factor of $\Theta$ is that corresponding to 2+2 particles. However by the conservation law we can construct from $f_{\Theta}$ the form factors of the left and right components of the energy momentum tensor $T$ and $\overline{T}$ whose lowest form factors are of the type $2n+0$ and $0+2m$ respectively. These lowest form factors must coincide with the form factors of pure $k=1$ WZNW model i.e. with those of $SU(2)$-invariant Thirring model. One easily finds that it implies: \begin{eqnarray} &&2\pi i\ res _{\beta_2'=\beta _1'+\pi i}\sum e^{-\beta _j}\ f(\beta_1,\cdots,\beta_{2n}\ |\ \beta_1',\beta_{2}') = \nonumber \\&&= \widehat{f}_{T}(\beta_1,\cdots,\beta_{2n})\otimes c_{1',2'} \(1- S_{LR}(\beta _1'-\beta_1)\cdots S_{LR}(\beta _1'-\beta_1)\) , \nonumber \\ && 2\pi i \ res _{\beta_2=\beta _1+\pi i}\sum e^{\beta _j'}\ f(\beta_1,\beta_{2}\ |\ \beta_1',\cdots,\beta_{2m}')= \nonumber \\&&= \widehat{f}_{\overline{T}}(\beta_1',\cdots,\beta_{2m}') \otimes c_{1',2'} \(1- S_{RL}(\beta _1-\beta_1')\cdots S_{RL}(\beta _1-\beta_1')\) \label{c2} \end{eqnarray} where $\widehat{f}_{T}$ and $\widehat{f}_{\overline{T}}$ are the form factors of left and right components of the energy-momentum tensor for the $SU(2)$ -Thirring model \cite{book}. 3. The IR limit corresponds to $\beta _i-\beta _j '\simeq\log\Lambda$ and $\Lambda\to\infty$. In this limit one has to reproduce the operator $T\overline{T}$ which defines the direction of the flow in the IR region. So, we must have \begin{eqnarray} f_{\Theta}(\beta_1+\log\Lambda,\cdots,\beta_{2n}+\log\Lambda|\ \beta_1',\cdots,\beta_{2m}') \to (M\Lambda) ^{-2 } \widehat{f}_{T}(\beta_1,\cdots,\beta_{2n}) \widehat{f}_{\overline{T}}(\beta_1',\cdots,\beta_{2m}') \label{c3} \end{eqnarray} Let us try to satisfy all this requirement. The simple form of the left-right S-matrix allows to exclude it from the equations (\ref{a1},\ref{a2}). Consider the function $g$ defined as follows: \begin{eqnarray} f_{\Theta}(\beta_1,\cdots,\beta_{2n}\ |\ \beta_1',\cdots,\beta_{2m}') =\prod\psi(\beta_i,\beta_j')g(\beta_1,\cdots,\beta_{2n}\ |\ \beta_1',\cdots,\beta_{2m}') \label{fg} \end{eqnarray} where $$\psi(\beta,\beta ')= 2^{-{3\over 4}} \exp \( -{1\over 4}(\beta+\beta ') -\int _0^{\infty} {2\sin ^2{1\over 2}(\beta-\beta ' +\pi i)k+\sinh ^2 {\pi k\over 2} \over 2k \sinh \pi k \cosh {\pi k\over 2}}dk\)$$ The function $\psi (\beta,\beta ')$ satisfies the equations \begin{eqnarray} &&\psi (\beta,\beta '+2\pi i)=\psi (\beta,\beta ')S_{RL}(\beta '-\beta),\quad \psi (\beta+2\pi i,\beta ')=\psi (\beta,\beta ')S_{LR}(\beta -\beta ') \nonumber \\ && \psi (\beta,\beta '+\pi i) \psi (\beta,\beta ')={1\over e^{\beta}-i e^{\beta '}}, \quad \psi (\beta +\pi i,\beta ') \psi (\beta,\beta ')={1\over ie^{\beta}- e^{\beta '}} \label{psi} \end{eqnarray} It is clear that the equation (\ref{a2}) rewritten in terms of $g$ does not contain the left-right S-matrices which means that the function $g$ must be of the form \begin{eqnarray} g(\beta_1,\cdots,\beta_{2n}\ |\ \beta_1',\cdots,\beta_{2m}')=\sum\limits _{K,L} c _{K,L}(\beta_1,\cdots,\beta_{2n}\ |\ \beta_1',\cdots,\beta_{2m}') \widehat{f}^K(\beta_1,\cdots,\beta_{2n})\widehat{f}^L (\beta_1',\cdots,\beta_{2m}') \label{xx} \end{eqnarray} where $\widehat{f}^K(\beta_1,\cdots,\beta_{2n})$ and $\widehat{f}^L (\beta_1',\cdots,\beta_{2m}')$ are different singlet solutions (counted by $K$ and $L$ whose nature will be explained later) of the equations \begin{eqnarray} &&\widehat{f}^K(\cdots,\beta_i,\beta_{i+1},\cdots)S^Y(\beta_i-\beta_{i+1} )= \widehat{f}^K(\cdots,\beta_{i+1},\beta_i,\cdots), \nonumber \\ && \widehat{f}^K(\beta_1,\cdots,\beta_{2n-1},\beta_{2n}+2\pi i)= \widehat{f}^K(\beta_{2n},\beta_1,\cdots,\beta_{2n-1}), \label{leq}\\ && \widehat{f}^L(\cdots,\beta_i',\beta_{i+1}',\cdots)S^Y(\beta_i'-\beta_{i+1}' )= \widehat{f}^L(\cdots,\beta_{i+1}',\beta_i',\cdots), \nonumber \\ && \widehat{f}^L(\beta_1',\cdots,\beta_{2m-1}',\beta_{2m}'+2\pi i)= \widehat{f}^L(\beta_{2m}',\beta_1',\cdots,\beta_{2m-1}'), \nonumber \end{eqnarray} The functions $c _{K,L}(\beta_1,\cdots,\beta_{2n}\ |\ \beta_1',\cdots,\beta_{2m}') $ are quasiconstants: $2\pi i$-periodical symmetrical with respect to $ \beta_1,\cdots,\beta_{2n} $ and $ \beta_1',\cdots,\beta_{2m}'$ functions with possible singularities only at $\beta_i,\beta_i'=\pm\infty$. The equations for left and right parts are the same, so, let us concentrate for the moment only on the left one. It is well known \cite{book,count} that the solutions to the equations (\ref{leq}) are counted by the functions $K(A_1,\cdots , $ $A_{n-1}| B_1,\cdots ,B_{2n})$ which are antisymmetrical polynomials of $A_1,\cdots ,A_{n-1}$ such that $1\le deg_{A_i}(K)\le 2n-1$, $\forall i$ and symmetrical Laurent polynomials of $B_1,\cdots ,B_{2n} $. The solutions are given by the formula \begin{eqnarray} &&\widehat{f}^K(\beta_1,\cdots ,\beta_{2n})=d^n \exp\({n\over 2}\sum \beta _j\)\prod\limits _{i,j}\zeta(\beta _i-\beta _j) \label{int} \\&& \times\int\limits _{-\infty}^{\infty}d\alpha _1\cdots \int\limits _{-\infty}^{\infty}d\alpha _{n-1} \prod\limits_{i,j} \widetilde{\varphi}(\alpha _i,\beta _j) \langle\Delta _n^{(0)}\rangle _n (\alpha _1,\cdots ,\alpha _{n-1} |\beta _{1},\cdots ,\beta _{2n} ) K(e^{\alpha _1},\cdots ,e^{\alpha _{n-1}} |e^{\beta _{1}},\cdots ,e^{\beta _{2n}}) \nonumber \end{eqnarray} where $$ \widetilde{\varphi}(\alpha _i,\beta _j) =e^{-{1\over 2}(\alpha +\beta)} \Gamma \({1\over 4}+{\alpha -\beta\over 2\pi i} \) \Gamma \({1\over 4}-{\alpha -\beta\over 2\pi i} \) $$ We do not give here the formulae for $\langle\Delta _n^{(0)}\rangle _n (\alpha _1,\cdots ,\alpha _{n-1} |\beta _{1},\cdots ,\beta _{2n} ) $ which is a rational function of all variables with values in the tensor product of the isotopic spaces, for $\zeta (\beta)$ which is certain transcendental function and for the constant $d$: these formulae can be found in the book \cite{book} (Chapter 7). It has to be emphasized that the integral (\ref{int}) vanishes for two kinds of function $K$ \cite{count,bbs2}: \begin{eqnarray} &&K(A_1,\cdots ,A_{n-1}|B_1,\cdots ,B_{2n})= \sum\limits _{k=1}^{n-1}(-1)^k (P(A_k)-P(-A_k)) K'(A_1,\cdots ,\widehat{A_k},\cdots ,A_{n-1}|B_1,\cdots ,B_{2n}) , \nonumber \\&& K(A_1,\cdots ,A_{n-1}|B_1,\cdots ,B_{2n})= \sum\limits _{k<l}(-1)^{k+l} C(A_k,A_l) K''(A_1,\cdots ,\widehat{A_k},\cdots ,\widehat{A_l}\cdots ,A_{n-1} |B_1,\cdots ,B_{2n})\quad \quad \label{zero} \end{eqnarray} where $K'$,$K''$ are some polynomials of the less number of variables with the same properties as $K$, $P(A)=\prod\limits _j (A_k+iB_j)$ and $$ C(A_1,A_2 )={1\over A_1A_2}\left\{ {A_1-A_2\over A_1+A_2 } (P(A_1)P(A_2)-P(-A_1)P(-A_2)) + (P(-A_1)P(A_2)-P(A_1)P(-A_2))\right\} \label{C} $$ So, the polynomials $K$ are defined modulo the polynomials of the kind (\ref{zero}) , the fact that has been used in \cite{count} to show that we have correct number of solutions to (\ref{leq}). Combining (\ref{fg}),(\ref{xx}) and (\ref{int}) we find that the from factors satisfying (\ref{a1}) and (\ref{a2}) are of the form \begin{eqnarray} &&f_{\Theta}(\beta_1,\cdots,\beta_{2n}\ |\ \beta_1',\cdots,\beta_{2m}') =M^2\prod\psi(\beta_i,\beta_j') \prod\limits _{i,j}\zeta(\beta _i-\beta _j) \prod\limits _{i,j}\zeta(\beta _i'-\beta _j') \nonumber \\ &&\times \int\limits _{-\infty}^{\infty}d\alpha _1\cdots \int\limits _{-\infty}^{\infty}d\alpha _{n-1} \int\limits _{-\infty}^{\infty}d\alpha '_1\cdots \int\limits _{-\infty}^{\infty}d\alpha '_{m-1} \prod\limits _{i=1}^{n-1}\prod\limits_{j=1}^{2n} \widetilde{\varphi}(\alpha _i,\beta _j) \prod\limits _{i=1}^{m-1}\prod\limits_{j=1}^{2m} \widetilde{\varphi}(\alpha '_i,\beta '_j) \nonumber \\ &&\times \langle\Delta _n^{(0)}\rangle _n (\alpha _1,\cdots ,\alpha _{n-1} |\beta _{1},\cdots ,\beta _{2n} ) \langle\Delta _n^{(0)}\rangle _n (\alpha' _1,\cdots ,\alpha _{m-1}' |\beta _{1}',\cdots ,\beta _{2m}' ) \nonumber \\ &&\times M_{n,m}(e^{\alpha _1},\cdots ,e^{\alpha _{n-1}} | e^{\alpha _1'},\cdots ,e^{\alpha _{m-1}'}| e^{\beta _{1}},\cdots ,e^{\beta _{2n}}| e^{\beta _{1}'},\cdots ,e^{\beta _{2m}'}) \label{ff} \end{eqnarray} where $ M_{n,m}(A _1,\cdots ,A _{n-1} | A _1',\cdots ,A _{m-1}'| B _{1},\cdots ,B _{2n}| B _{1}',\cdots ,B _{2m}') $ is an antisymmetrical polynomial of $ A _1,$ $\cdots ,A _{n-1} $ ($A _1',\cdots ,A _{m-1}'$) whose degree with respect to every variable is from $1$ to $2n-1$ (from $1$ to $2m-1$) and symmetrical Laurent polynomial of $B _{1},\cdots ,B _{2n}$ ($B _{1}',\cdots ,B _{2m}'$). Now we have to satisfy the rest of our requirements on the form factors. In the paper \cite{count} there is a general prescription for handling the residue condition (\ref{a3}) for the integrals of the form (\ref{int}). Applying this prescription to our situation and using the equations (\ref{psi}) one finds that the residue condition (\ref{a3}) is satisfied if and only if the function $M_{n,m}$ possesses the properties: \newline First, \begin{eqnarray} &&M_{n,m}(A _1,\cdots ,A _{n-1} |A _1',\cdots ,A _{m-1}'| B _{1},\cdots ,B _{2n-2},B,-B|B _{1}',\cdots ,B _{2m}')= \nonumber \\&&= \sum\limits _{k=1}^{n-1}(-1)^k\prod\limits _{p\ne k}(A_p^2+B^2) M_{n-1,m}^k(A _1,\cdots ,A _{n-1} |A _1',\cdots ,A _{m-1}'| B _{1},\cdots ,B _{2n-2}|B|B _{1}',\cdots ,B _{2m}') \nonumber \\ && M_{n,m}(A _1,\cdots ,A _{n-1} |A _1',\cdots ,A _{m-1}'| B _{1},\cdots ,B _{2n}|B _{1}',\cdots ,B _{2m-2}',B',-B')= \label{rel1}\\&&= \sum\limits _{k=1}^{m-1}(-1)^k\prod\limits _{p\ne k}((A_p')^2+(B')^2) M_{n,m-1}^k(A _1,\cdots ,A _{n-1} |A _1',\cdots ,A _{m-1}'| B _{1},\cdots ,B _{2n}|B _{1}',\cdots ,B _{2m-2}'|B') \nonumber \end{eqnarray} where $M_{n-1,m}^k $ and $M_{n,m-1}^k$ are some {\it polynomials} in $A_i$ and $A_i'$. \newline Second, \begin{eqnarray} &&M_{n-1,m}^k(A _1,\cdots ,A _{k-1},\pm iB ,A _{k-1},\cdots ,A _{n-1} |A _1',\cdots ,A _{m-1}'| B _{1},\cdots ,B _{2n-2}|B|B _{1}',\cdots ,B _{2m}')= \nonumber \\&& = \pm B\prod\limits _{j=1}^{2m}(B\mp iB'_j) M_{n-1,m}(A _1,\cdots ,A _{k-1},A _{k-1},\cdots ,A _{n-1} |A _1',\cdots ,A _{m-1}'| B _{1},\cdots ,B _{2n-2}|B _{1}',\cdots ,B _{2m}'), \nonumber \\&& M_{n,m-1}^k(A _1,\cdots ,A _{n-1} |A _1',\cdots ,A _{k-1}',\pm iB' ,A _{k-1}',\cdots ,A _{m-1}'| B _{1},\cdots ,B _{2n}|B _{1}',\cdots ,B _{2m-2}'|B')= \label{rel2} \\ &&= \pm B'\prod\limits _{j=1}^{2m}(B'\pm iB_j) M_{n,m-1}(A _1,\cdots ,A _{n-1} |A _1',\cdots ,A _{k-1}',A _{k-1}',\cdots ,A _{m-1}'| B _{1},\cdots ,B _{2n}|B _{1}',\cdots ,B _{2m-2}') \nonumber \end{eqnarray} These equations are necessary and sufficient for the formula (\ref{ff}) to define form factors of a local operator. Certainly, they have infinitely many solutions. We shall give only one of these solutions corresponding to the operator $\Theta$. Let us introduce the notations for the sets of integers: $S=\{1,\cdots ,2n\}$, $S'=\{1,\cdots ,2m\}$ then \begin{eqnarray} &&M_{n,m}(A _1,\cdots ,A _{n-1} |A _1',\cdots ,A _{m-1}'| B _{1},\cdots ,B _{2n}|B _{1}',\cdots ,B _{2m}') = \prod\limits _{i<j}(A_i-A_j)\prod\limits _{i<j}(A_i'-A_j') \nonumber \\&& \times \prod\limits _{j=1}^{2n} B_j^{-1} \prod\limits _{i=1}^{n-1} A_i^2 \prod\limits _{i=1}^{m-1} A_i ' \sum\limits _{{T\subset S\atop\# T=n-1}} \sum\limits _{{T'\subset S'\atop\# T'=m-1}} \prod\limits _{j\in T} B_j \prod\limits _{i=1}^{n-1} \prod\limits _{j\in T}(A_i+iB_j) \prod\limits _{i=1}^{m-1} \prod\limits _{j\in T'}(A_i'+iB_j') \nonumber \\&& \times \prod\limits _{{i,j\in \overline{T}\atop i<j}}(B_i+B_j) \prod\limits _{{i,j\in \overline{T}'\atop i<j}}(B_i'+B_j') \prod\limits _{{i\in T\atop j\in \overline{T}}}{1\over B_i-B_j} \prod\limits _{{i\in T'\atop j\in \overline{T}'}}{1\over B_i'-B_j'} \nonumber \\&& \times \prod\limits _{{i\in T\atop j\in T'}}(B_i+iB_j') \prod\limits _{{i\in \overline{T}\atop j\in \overline{T}'}}(B_i-iB_j') \ X_{T,T'}(B_1,\cdots ,B_{2n}|B_1',\cdots ,B_{2m}') \label{poly} \end{eqnarray} where $\overline{T}=S\backslash T$, $\overline{T'}=S'\backslash T'$, \begin{eqnarray} X_{T,T'}(B_1,\cdots ,B_{2n}|B_1',\cdots ,B_{2m}')= \sum\limits _{i_1,i_2\in\overline{T}} \prod\limits _{p=1}^2\( {\prod\limits _{j\in T}(B_{i_p}+B_j) \over \prod\limits _{j\in \overline{T}\backslash \{i_1,i_2\}}(B_{i_p}-B_j)} {\prod\limits _{j\in T'}(B_{i_p}+iB_j') \over \prod\limits _{j\in \overline{T}'}(B_{i_p}-iB_j')} \) \nonumber \end{eqnarray} The polynomial $X_{T,T'}(B_1,\cdots ,B_{2n}|B_1',\cdots ,B_{2m}') $ is in fact quite symmetric with respect to replacement $B\leftrightarrow B'$. Let us show that $M_{n,m}$ satisfies all necessary requirements. The relations (\ref{rel1},\ref{rel2}) are easily checked using the formula $$\prod\limits _{i=1}^{n-1} \prod\limits _{j=1}^{n-1}(A_i+iB_j) \prod\limits _{i<j}(A_i-A_j)= \prod\limits _{i=1}^{n-1} \prod\limits _{j=1}^{n-1}(A_i^2+B_j^2) \prod\limits _{i<j}^{n-1}{1\over B_i-B_j}\ det\({1\over A_i-iB_j}\) $$ So, $M_{n,m}$ really defines a local operators. We have to show that the additional conditions formulated at the beginning of this section are satisfied in order to show that this local operator is indeed the trace of the energy-momentum tensor. Obviously, $M_{n,m}$ is a homogeneous function of all its variables ($A,B,A',B'$) of total degree $(m+n)^2-2m-2n$. Considering the formula (\ref{ff}) one realizes that this fact provides that the operator defined by $M_{n,m}$ is Lorentz scalar i.e. its form factors are invariant under simultaneous shift of all the rapidities. Let us consider now the conditions 1-3 formulated earlier. We start form the condition 3. One finds that $$\psi (\beta +\log \Lambda ,\beta ')\to \Lambda ^{-{1\over 2}} e^{-{1\over 2}\beta} $$ The integrals with respect to $\alpha _i$ in (\ref{ff}) are concentrated near the points $\beta _j$, so when $\beta_j$ become of order $\log \Lambda $ the integration variables $\alpha _i$ must be of the same order. One finds that when $\log \Lambda \to \infty$ \begin{eqnarray} &&M_{n,m}(\Lambda A _1,\cdots ,\Lambda A _{n-1} |A _1',\cdots ,A _{m-1}'| \Lambda B _{1},\cdots ,\Lambda B _{2n}|B _{1}',\cdots ,B _{2m}') \to \nonumber \\ &&\to \Lambda ^{2mn+n^2-2n-2} \prod\limits _{j=1}^{2n} B_j^{2m-1} \sum\limits _{j=1}^{2n} B_j^{-1} \prod\limits _{i=1}^{n-1} A_i^3 \prod\limits _{i<j}(A_i^2-A_j^2) \sum\limits _{j=1}^{2m} B_j ' \prod\limits _{i=1}^{m-1} A_i ' \prod\limits _{i<j}(A_i'^2-A_j'^2) \nonumber \end{eqnarray} This formula is equivalent to (\ref{c3}) because the form factors of the energy-momentum tensor of $SU(2)$-invariant Thirring model $\widehat{f}_{T}(\beta_1,\cdots,\beta_{2n})$ $\widehat{f}_{\overline{T}}(\beta_1',\cdots,\beta_{2m}') $ are given by the formulae on the type (\ref{int}) with the polynomial $K$ equal respectively to \begin{eqnarray} M^2\prod\limits _{j=1}^{2n} B_j^{-1} \sum\limits _{j=1}^{2n} B_j^{-1}\prod\limits _{i=1}^{n-1} A_i^3 \prod\limits _{i<j}(A_i^2-A_j^2) \qquad and \qquad M^2\sum\limits _{j=1}^{2m} B_j ' \prod\limits _{i=1}^{m-1} A_i ' \prod\limits _{i<j}(A_i'^2-A_j'^2) \label{T} \end{eqnarray} Let us consider the condition (\ref{c2}). One finds that \begin{eqnarray} &&\left. {1\over B _{1}' +B _{2}' } M_{n,1}(A _1,\cdots ,A _{n-1} |\emptyset| B _{1},\cdots ,B _{2n}|B _{1}' ,B _{2}') \right| _{B _{2}' =-B _{1}'}= \nonumber \\&&\hskip 1cm = \prod\limits _{j=1}^{2n} B_j^{-1} \prod\limits _{i=1}^{n-1} A_i^3 \prod\limits _{i<j}(A_i^2-A_j^2) \(\prod (B_1'+iB_j )-\prod (B_1'-iB_j )\) \nonumber \end{eqnarray} which together with (\ref{T}) gives the first equation from (\ref{c2}), the second relation is proven similarly. The condition (\ref{c1}) is the most complicated to prove. Naively it has to be equivalent to the fact that $M_{n,m} $ is divisible by $\sum B_j^{-1}$ and $\sum B_j'$, but that is not the case: the function $M_{n,m} $ has to be substituted into the integral hence it is defined modulo the functions of the type (\ref{zero}) (and similar functions of $A_i'$). Thus the divisibility has to be proven modulo these null-polynomials. We have checked this fact for many particular examples, but still we lack a general proof. However, the calculations in particular cases go so nicely that we have no doubt that the relation (\ref{c1}) is satisfied generally. \section{Some generalizations.} The model considered in this paper provides a special case of wide class of massless flows. Consider the massless flow \cite{sot} between the UV coset model $su(2)_{k+1}\otimes su(2)_k/su(2)_{2k+1}$ and the IR coset model $su(2)_{k}\otimes su(2)_1/su(2)_{k+1}$, the latter model is nothing but the minimal model $M_{k+2}$. This flow is defined in UV by the relevant operator of dimension $1-2/(2k+3) $, it arrives at IR region along $T\overline{T}$. The massless S-matrices for these flows are written in terms of RSOS restriction of the sine-Gordon (SG) S-matrix $S^{\xi}(\beta)$ ($\xi$ is SG coupling constant defined as in \cite{book}). Namely \cite{bl}, $$S_{LL}(\beta _1,\beta_2)=S^{\pi(k+2)}_{RSOS}(\beta _1-\beta_2),\qquad S_{RR}(\beta _1',\beta_2')=S^{\pi(k+2)}_{RSOS}(\beta _1'-\beta_2') $$ The left-right S-matrix is independent of $k$, it is the same as above. When $k=\infty$ the model coincides with PCM${}_1$. Another extreme case is $k=1$ when the model describes the flow between tricritical and critical Ising models. It is well known that the RSOS-restriction for $\xi =3\pi$ effectively reduces soliton to one-component particle with free scattering: $$S^{\pi(k+2)}_{RSOS}(\beta)=-1$$ The results of this paper allow straightforward generalization to these flows. One has to replace the formulae of the type (\ref{int}) by their SG analogues. This does not disturb the function $M_{n,m}$ because the way of counting solution to the equation of the type (\ref{leq}) in SG case does not depend on the coupling constant as well as all the equations on $M_{n,m}$. So, the form factors are defined by (\ref{ff}) where one has to replace the functions $\zeta$, $\varphi$, $\langle \Delta ^{(0)}\rangle$ by their SG-analogues and to take RSOS restriction. Let us see how it works in the case $k=1$. For generic coupling constant the formulae (\ref{zero}) present the only reason for vanishing the integrals of the type (\ref{int}). However when $ \xi=3\pi$ and RSOS restriction is taken the integral does not vanish only for the antisymmetrical polynomial $K(A_1,\cdots ,A_{n-1})$ of very special kind: $$ K(A_1,\cdots ,A_{n-1}) =\prod A_i^2\prod\limits _{i<j}(A_i^2-A_j^2)$$ The value of the integral for this kind of polynomial (taking in account the functions $\zeta$ also) is $$\prod\limits _{i<j}\tanh {1\over 2}(\beta_i-\beta _j) \exp\({1\over 2}\sum \beta_j \)$$ Consider the formula (\ref{poly}). We have to take the functions $$ \prod\limits _{i<j}(A_i-A_j) \prod\limits _{i=1}^{n-1} A_i^2 \prod\limits _{i=1}^{n-1} \prod\limits _{j\in T}(A_i+iB_j) \quad and\quad \prod\limits _{i<j}(A_i'-A_j') \prod\limits _{i=1}^{m-1} A_i ' \prod\limits _{i=1}^{m-1} \prod\limits _{j\in T'}(A_i'+iB_j'), $$ to decompose them with respect to antisymmetrical polynomials of $A_i$ and $A'_i$ corresponding to different partitions and to find the coefficients with which enter the polynomials $\prod A_i^2\prod\limits _{i<j}(A_i^2-A_j^2)$ and $\prod (A'_i)^2\prod\limits _{i<j}((A'_i)^2-(A'_j)^2)$. These coefficients are $$ \prod\limits _{j\in T}B_j \prod\limits _{{i,j\in T\atop i<j}}(B_i+B_j) \quad and \quad \prod\limits _{{i,j\in T'\atop i<j}}(B_i'+B_j') $$ Thus we find the following formula for the form factors of $\Theta$ for this model \begin{eqnarray} &&f_{\Theta}(\beta_1,\cdots,\beta_{2n}\ |\ \beta_1',\cdots,\beta_{2m}') =M^2 \prod\psi(\beta_i,\beta_j') \prod\limits _{i,j}\tanh {1\over 2}(\beta_i-\beta _j) \prod\limits _{i,j}\tanh {1\over 2}(\beta_i-\beta _j) \nonumber \\ &&\times \exp\({1\over 2}\sum \beta_j+{1\over 2}\sum \beta_j' \) Q_{n,m} (e^{\beta _{1}},\cdots ,e^{\beta _{2n}}| e^{\beta _{1}'},\cdots ,e^{\beta _{2m}'}) \nonumber \end{eqnarray} where \begin{eqnarray} &&Q_{n,m}( B _{1},\cdots ,B _{2n}|B _{1}',\cdots ,B _{2m}') = \nonumber \\&& = \prod\limits _{j=1}^{2n} B_j^{-1} \sum\limits _{{T\subset S\atop\# T=n-1}} \sum\limits _{{T'\subset S'\atop\# T'=m-1}} \prod\limits _{j\in T} B_j^2 \prod\limits _{{i,j\in T\atop i<j}}(B_i+B_j) \prod\limits _{{i,j\in T'\atop i<j}}(B_i'+B_j') \nonumber \\&& \times \prod\limits _{{i,j\in \overline{T}\atop i<j}}(B_i+B_j) \prod\limits _{{i,j\in \overline{T}'\atop i<j}}(B_i'+B_j') \prod\limits _{{i\in T\atop j\in \overline{T}}}{1\over B_i-B_j} \prod\limits _{{i\in T'\atop j\in \overline{T}'}}{1\over B_i'-B_j'} \nonumber \\&& \times \prod\limits _{{i\in T\atop j\in T'}}(B_i+iB_j') \prod\limits _{{i\in \overline{T}\atop j\in \overline{T}'}}(B_i-iB_j') X_{T,T'}(B_1,\cdots ,B_{2n}|B_1',\cdots ,B_{2m}') \label{kon} \end{eqnarray} one can write a formula for this polynomial in determinant form, but we think that (\ref{kon}) shows quite transparently how all the required properties of this polynomial \cite{muss} are satisfied.
proofpile-arXiv_065-459
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{INTRODUCTION} \begin{figure} \psfig{file=snfig.ps,height=9.53cm,rheight=8.3cm} \caption{\label{snfig}\small Direct-capture cross sections at 30 keV for different Sn isotopes. Levels and masses are calculated with models by Sharma et al.~\protect{\cite{sharma}} (triangles), M\"oller et al.~\protect{\cite{moeller}} (dots), and Dobaczewski et al.~\protect{\cite{doba}} (squares). The lines are drawn to guide the eye.} \end{figure} \begin{figure} \psfig{file=nic96figc.ps,height=9.53cm,rheight=8.3cm} \caption{\label{states}\small Dependence of level energies on mass number for the even-odd isotopes $^{125-135}$Sn in the RMFT model~\protect{\cite{sharma}} (right) and for the isotopes $^{125-145}$Sn in the HFB model~\protect{\cite{doba}} (left). Shown are the 1/2$^-$ state (open circles), the 3/2$^-$ state (triangles) and the calculated neutron separation energy (full circles). The lines are drawn to guide the eye. Note the different range in mass numbers in the two plots.} \end{figure} Explosive nuclear burning in astrophysical environments produces unstable nuclei which again can be targets for subsequent reactions. Most of these nuclei are not accessible in terrestrial labs or not fully explored by experiments, yet. For the majority of unstable nuclei the statistical model (Hauser-Feshbach) can be used to determine the cross sections. However, for nuclei close to the dripline the level density becomes too low to apply the statistical model~\cite{tommy} and contributions of a direct interaction mechanism (DI) may dominate the cross sections. The DI requires the detailed knowledge of energy levels (excitation energies, spins, parities), contrary to the statistical model which averages over resonances. Lacking experimental data, this information has to be extracted from microscopic nuclear-structure models. We compare the results for direct neutron capture (calculated in the optical model~\cite{kim}) on the even-even isotopes $^{124-145}$Sn with energy levels, masses, and nuclear density distributions taken from different nuclear-structure models. The utilized structure models were a Hartree-Fock-Bogoliubov model (HFB) with SkP force~\cite{doba}, a Relativistic Mean Field Theory (RMFT) with the parameter set NLSH~\cite{sharma} and a Shell Model based on folded-Yukawa wave functions (FYSM)~\cite{moeller}. A similar study has already been performed for neutron-rich Pb isotopes~\cite{tom}. \section{METHOD} The cross sections were calculated in the optical model for direct capture~\cite{kim}, utilizing optical potentials derived by the folding procedure~\cite{satch}. In the folding approach the nuclear target density is folded with an energy- and density-dependent effective nucleon-nucleon interaction in order to obtain the potentials for the bound and scattering states. Only one open parameter $\lambda$ remains which accounts for the effects of antisymmetrization and is close to unity. The densities required for the determination of the folding potentials were consistently calculated from the wave functions of the respective nuclear-structure model. For the bound states the strength parameter $\lambda$ was fixed by the condition to reproduce the given binding energy of the captured neutron. The value of $\lambda$ for the scattering potential was adjusted to yield the same value of 425 MeV fm$^3$~\cite{werner} for the volume integral as determined from the experimental scattering data on stable Sn isotopes~\cite{mug,cinda}. In order to be able to directly compare the different models, all nuclei were assumed to be spherical and the spectroscopic factors were set to 1. \section{RESULTS AND DISCUSSION} The results of the calculations for projectiles at $E_{\mathrm{c.m.}}=30$ keV are summarized in Fig.~\ref{snfig}. For each model we calculated the capture cross section only up to the r-process path. The most extreme location of the path (farthest away from the line of stability) is determined by neutron separation energies $E_{\mathrm{n}} \approx 2$ MeV~\cite{fkt}. Depending on the microscopic model, the path will then be located at higher or lower mass numbers $A$. In the case of RMFT and FYSM it will go through $A\approx$ 132--134, for HFB the path will be shifted to considerably higher mass numbers $A\approx$ 142--144. (The neutron dripline is also shifted to higher masses in the latter model.) Similar effects as seen in the behavior of the Pb cross sections~\cite{tom} can also be found for the Sn cross sections. The cross section can vary by order of magnitudes when going from one isotope to the next and also differ vastly between the different microscopic models. As the capture to low-spin states ($J$=1/2, 3/2) accounts for the largest contributions to the cross section, the results are very sensitive to the presence of bound low-spin states. Since the microscopic models not only yield different masses (i.e.\ neutron separation energies) but also exhibit different behaviors of the level energies with changing mass, ``jumps'' and ``gaps'' can be seen with some models (RMFT, FYSM), whereas others (HFB) result in a smoother behavior of the capture cross sections in an isotopic chain. This is illustrated in Fig.~\ref{states}, which shows the neutron separation energy and the excitation energy of the 1/2$^-$ and 3/2$^-$ states in RMFT and HFB. As long as both states are unbound in the RMFT, the cross sections remain low and only jump to higher values when those states become bound at the shell closure. As at least the 3/2$^-$ level is always bound in HFB, the cross sections show a smoother behavior. The variation in the FYSM cross sections can be explained in a similar way. \section{CONCLUSION} With this work we have underlined that the calculation of purely theoretical direct capture cross sections far from stability still contains a large error, even when using most recent nuclear-structure models. In the previously discussed case of Pb isotopes~\cite{tom}, the r-process path contains nuclei in or at the border of a region expected to be deformed, leading to higher level densities and thus favoring the compound nucleus mechanism. This is not true for neutron-rich isotopes in the Sn region, especially around the neutron magic number $N=82$ where the level density becomes too low for the statistical model. Therefore the neutron capture cross sections have to be calculated using input from nuclear-structure models and will be subject to the quoted uncertainties, even when the different models yield similar values for other nuclear properties, such as masses. Similar problems may be encountered on the proton-rich side when predicting proton capture cross sections close to the proton dripline.
proofpile-arXiv_065-460
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} {\Large 1. Introduction}\\ \noindent One of the most precise and from the theoretical point unambiguous determinations of the strong coupling constant $\alpha_s$ has been obtained from the hadronic decay rate of the $Z$ boson. Ingredients are the large event rate which allows to reduce the statistical uncertainty in $\alpha_s$ down to $0.003$, the precise calibration through the accurate measurement of the luminosity or the leptonic rate \cite{Blondel} and, last but not least, the elaborate theoretical calculations \cite{YellowReport,CKKRep}. This high energy determination of $\alpha_s$ should and could be complemented by a similar measurement of $R_{\rm had}$ in the energy region just below the $B$ meson threshold: large event rates are available at the $e^+ e^-$ storage ring CESR and an experimental analysis is currently underway \cite{DaveBessonpriv}. Theoretical predictions for $R_{\rm had}$ below the $B{\overline B}$ threshold have been presented earlier \cite{rhad}. These include terms up to order $\alpha_s^3$ for the massless limit and for $m_c^2/s$ corrections as well, terms of order $\alpha_s^2$ for the quartic mass terms $(m_c^2/s)^2$, terms of order $\alpha_s (m_c^2/s)^3$ and contributions from virtual bottom quarks up to order $\alpha_s^2 (s/m_b^2)$. Recently it has been demonstrated that this expansion in $(m_c^2/s)$ provides an excellent approximation not only around $\sqrt{s} = 10$~GeV, but also down to 6 or even 5 GeV. \par The forthcoming experimental analysis necessitates on the one hand a thorough understanding of the background, e.g. from two photon events or the feedthrough from $\tau$ pairs into the hadronic cross section, and, on the other hand, a realistic prediction for the annihilation channel. The former is outside the scope of this paper, the latter is the subject of this work. It requires to incorporate the effect of the running QED coupling constant and initial state radiation. The photon vacuum polarization leads to an increase of the cross section by about 7--8\%. Magnitude and even sign of initial state corrections depend on the experimental procedure, in particular on the minimal mass of the hadronic system accepted for the event sample. Another important issue is the treatment of the tails of the $\Upsilon$ resonances, which contribute coherently through mixing with the photon and incoherently through their radiative tails. In order to study these effects and the associated theoretical uncertainties separately the corresponding analysis for the reaction $e^+ e^- \to \tau^+ \tau^-$ is presented in section 2. These results may also serve as an independent experimental calibration of the cross section. The corresponding predictions for the hadronic cross section are discussed in section 3. In section 4 we comment on the accuracy of the presented approach and section 5 contains our summary and conclusions. \vskip 1 cm \noindent {\Large 2. Lepton Pair Production}\\ \noindent To calibrate the predictions for the hadron production cross section it seems appropriate and useful to calculate in a first step the cross section for lepton pair production. To arrive at a reliable prediction, initial and final state radiation, the effect of vacuum polarization and the influence of the nearby $\Upsilon$ resonances must be included. In this section results are presented for $\tau$ pair production. \vskip 0.5 cm \noindent {\bf (a) Initial state radiation} \noindent The most important correction to the total cross section is introduced by initial state radiation. It leads to a reduction of the invariant mass of the produced lepton pair or hadronic system and, for fairly loose cuts, to a significant enhancement of the cross section, albeit with events of significantly lower invariant mass of the system of interest. For the precise determination of $R_{\rm had}$ discussed below it is advisable to exclude the bulk of these low mass events. This reduces the size of the correction and at the same time the dependence on the input for $R_{\rm had}$ from the lower energy region. The treatment of initial state radiation has advanced significantly as a consequence of the detailed calculations performed for the analysis of the $Z$ line shape. The result for the inclusive cross section can be written in the form \begin{equation} \sigma(s) = \int_{z_0}^1 {\rm d}z\,\sigma_0(sz)\,G(z)\,. \label{eqisrint} \end{equation} The cross section including photon vacuum polarization (running $\alpha$) is denoted by $\sigma_0$. The invariant mass of the produced fermion pair is given by $s\,z$, where \begin{equation} \frac{m_{\rm min}^2}{s} \leq z_0 \leq z \leq 1 \label{eqisrborder} \end{equation} and $m_{\rm min} = 2 m_\ell$ for lepton pairs and $m_{\rm min} = 2 m_\pi$ for hadron production. In the cases of interest for this paper $z_0$ will have to be adopted to the experimental setup. Typically it is significantly larger than the theoretically allowed minimal value. For $E_{\rm cm} = 10.52$~GeV a cut in $z$ around $0.25$ corresponding to roughly $5$~GeV in the minimal mass will in the case of hadronic final states exclude the charmonium resonance region and the charm threshold region as well, thus limiting the hadron analysis to truely multihadronic final states. The complete radiator function $G(z)$ up to order $\alpha^2$ has been calculated in \cite{Burgers,vanNeerven}. The resummation of leading logarithms is discussed in \cite{YRBerends} (see also \cite{Kuraev}). For the present purpose an approximation is adequate which is exact in order $\alpha$ and which includes the dominant terms of order $\alpha^2$ plus leading logarithms. For the radiator function $G$ we thus take \cite{JadachWard} \begin{equation} G(z) = \beta (1-z)^{\beta-1}\,{\rm e}^{\delta_{yfs}}\,F\,\left( \delta_C^{V+S} + \delta_C^H \right)\,, \label{eqGc} \end{equation} with \begin{eqnarray} \beta &=& \frac{2\alpha}{\pi}(L-1)\,,\nonumber\\ L &=& \ln\frac{s}{m_e^2}\,,\nonumber\\ \delta_{yfs} &=& \frac{\alpha}{\pi} \left( \frac{L}{2} - 1 + 2\zeta(2) \right)\,,\nonumber\\ \delta_c^{V+S} &=& 1+\frac{\alpha}{\pi}(L-1)+\frac{1}{2} \left(\frac{\alpha}{\pi}\right)^2 L^2\,,\nonumber\\ \delta_C^H &=& -\frac{1-z^2}{2}+\frac{\alpha}{\pi}L\left[-\frac{1}{4} \left(1+3z^2\right)\ln z -1+z\right]\,,\nonumber\\ F &=& \frac{{\rm e}^{-\beta\gamma_E}}{\Gamma(1+\beta)}\,,\nonumber \end{eqnarray} where $\gamma_E = 0.5772\ldots$ is Euler's constant. Eq.~(\ref{eqGc}) is suited for quick numerical integration. The difference to the initial state radiation convolution using the complete order $\alpha^2$ result (eq~(3.12) of \cite{YRBerends}) is below one permille. The predictions for the reference energy of $10.52$~GeV and a variety of cuts $z_0$ are listed in Table~\ref{table1}. \begin{table}[htb] \caption{Dependence of the cross section $e^+ e^- \to \tau^+ \tau^-$ on the cutoff $z_0$ for $E_1 = 10.52$~GeV.} \begin{center} \begin{tabular}{|c||c|c|c|c|c|c|c|} \hline $z_0$ & 0.114 & 0.20 & 0.25 & 0.5 & 0.8 & 0.9 & 0.95 \\ \hline $m_{\rm cut}\ [GeV]$ & $2m_{\tau}$ & 4.705 & 5.260 & 7.439 & 9.409 & 9.980 & 10.254 \\ \hline\hline $\sigma\ [nb]$ & 0.9269 & 0.9101 & 0.8998 & 0.8543 & 0.7831 & 0.7359 & 0.6922 \\ \hline $\sigma/\sigma_{pt}$ & 1.1822 & 1.1608 & 1.1477 & 1.0896 & 0.9988 & 0.9386 & 0.8829 \\ \hline \end{tabular} \end{center} \label{table1} \end{table} \begin{figure}[htb] \begin{center} \leavevmode \epsfxsize=17.cm \epsffile[70 290 510 530]{taudiff.ps} \vskip -3mm \caption[]{\label{fig1ab} {\em (a) The differential cross section ${\rm d}\sigma(e^+ e^- \to \tau^+ \tau^-)/{\rm d}z$ and (b) the cutoff dependent cross section $\sigma(z > z_0)$ in nb for the centre of mass energy $E_1 = 10.52$~GeV as functions of $z$ and $z_0$, respectively.}} \end{center} \end{figure} The differential distribution ${\rm d}\sigma/{\rm d}z$ and the cutoff dependent cross section $\sigma(z > z_0)$ are displayed in Fig.~\ref{fig1ab}a, b, respectively, as functions of $z$, $z_0$. For the cross section before convolution the corrections from the vacuum polarization, the tails of the $\Upsilon$ resonances and final state radiation are included. Initial state radiation of lepton pairs and of hadrons has been treated in \cite{vanNeerven,KKKS}. These could easily be incorporated in the present formalism. The resulting corrections are at the permille level and will be ignored for the moment. Corrections from initial state radiation are evidently huge for loose or extremely tight cutoffs, with opposite sign. The strong cutoff dependence of the result is apparent from Table \ref{table1} and Fig.~\ref{fig1ab} as well. The result is relatively stable against variations of $z_0$ for values of $z_0$ below 0.75. At the same time the sensitivity to the input for $\sigma_0$ from smaller $s' = s\,z$ is eliminated for this choice. From this viewpoint a value around 0.7 to 0.8 seems optimal for the experimental determination of $R_{\rm had}$. On the other hand, if one allows to use input from smaller $s'$ from other experiments or from the measured events with photons from initial state radiation, then also lower values of $z_0$ down to about 0.25 are equally acceptable. The distribution ${\rm d}\sigma/{\rm d}z$ is flat in the region $0.25 < z < 0.75$ and numerically small. An experimental analysis based on a cutoff $z_0 = s'/s$ with a reasonably symmetric resolution in $s'$ is thus insensitive towards the details of the resolution function. In fact, even an uncontrolled shift of the central value by $\Delta z_0 = 0.05$ would lead to a 1\% deviation of the cross section only. \noindent {\bf (b) Final state radiation} \noindent Initial state radiation is strongly enhanced, a consequence of the inherently large logarithm in the correction function (cf. eq.~(\ref{eqGc})). Final state radiation with loose cuts, in contrast, is typically of order $\alpha/\pi$ without a large logarithm. In the massless limit and without a cutoff on the invariant fermion pair mass the correction factor $(1 + 3/4\,\alpha/\pi)$ amounts to two permille only. It depends, however, on the lepton mass and the cutoff. For the relatively loose cuts with $z_0$ between 0.25 and 0.8 one may well use the totally inclusive correction function $r^{(1)}$ defined by \begin{equation} R_{f\bar f}\, := \, \frac{\sigma(e^+e^-\to\gamma^*\to f\bar f\ldots)} {\sigma_{pt}} \quad = \quad r^{(0)} + \,\bigg(\frac{\alpha}{\pi}\bigg)\,r^{(1)} + \ldots\,, \label{eqrdef} \end{equation} with $\ \sigma_{pt} = 4\pi\alpha^2/3s\ $ and \begin{eqnarray} r^{(0)} & = & \frac{\beta}{2}\,(3-\beta^2)\,,\\[2mm] r^{(1)} & = & \frac{\left( 3 - {\beta^2} \right) \,\left( 1 + {\beta^2} \right) }{2 }\,\bigg[\, 2\,\mbox{Li}_2(p) + \mbox{Li}_2({p^2}) + \ln p\,\Big( 2\,\ln(1 - p) + \ln(1 + p) \Big) \,\bigg] \,\nonumber\,\\ & & \mbox{} - \beta\,( 3 - {\beta^2} ) \, \Big( 2\,\ln(1 - p) + \ln(1 + p) \Big) - \frac{\left( 1 - \beta \right) \, \left( 33 - 39\,\beta - 17\,{\beta^2} + 7\,{\beta^3} \right) }{16}\, \ln p\,\nonumber\,\\ & & \mbox{} + \frac{3\,\beta\,\left( 5 - 3\,{\beta^2} \right) }{8} \,, \end{eqnarray} where \begin{eqnarray} p \, = \, \frac{1-\beta}{1+\beta}\,,\qquad\, \beta \, = \, \sqrt{1-4m_f^2/s}\,. \end{eqnarray} In the case of hadron production, the main motivation of this investigation, photons from final state radiation will in general anyhow be included in the hadronic invariant mass. \vskip 0.5 cm \noindent {\bf (c) Leptonic and hadronic vacuum polarization} \noindent The leptonic vacuum polarization in one loop approximation is given by \begin{equation} \widehat\Pi_{\gamma\gamma}(s) := \Pi_{\gamma\gamma}(s) - \Pi_{\gamma\gamma}(0) = \frac{\alpha}{3\pi}\,\sum_f N_c \, Q_f^2 \, P(s, m_f)\,, \label{eqvacpollepdef} \end{equation} with \begin{equation} P(s, m_f) = \frac{1}{3} - \left(1+\frac{2m_f^2}{s}\right)\, \left(2+\beta\ln\frac{\beta-1}{\beta+1}\right)\,. \label{eqvacpollepp} \end{equation} For $|s| \gg m_{\ell}^2$ the function $P(s, m_f)$ is well approximated by \begin{equation} P_{\rm asymp.}(s, m_f) = -\frac{5}{3} + \ln\bigg(-\frac{s}{m_f^2} + i \,\epsilon\bigg)\,. \label{eqvacpolleppappr} \end{equation} The hadronic vacuum polarization is obtained from it's imaginary part via a dispersion relation \begin{equation} {\rm Re}\widehat\Pi_{\rm had}(q^2) = \frac{\alpha q^2}{3\pi} \, {\rm P} \int_{m_{\pi}^2}^{\infty} \frac{R_{\rm had}(s')}{s'(s'-q^2)} \, {\rm d}s'\,. \label{eqvacpolhad} \end{equation} To avoid complications that arise from the numerical evaluation of the dispersion integral over the data in the timelike region, the integral is evaluated at the corresponding value of $s$ in the spacelike region. We use parametrisations from the evaluation of eq.~(\ref{eqvacpolhad}) in the spacelike region as provided by \cite{jegerl,burkhardt}. The results for the complete cross sections calculated with the different parametrisations from \cite{jegerl} and \cite{burkhardt} agree to better than $10^{-4}$ in the region of interest. The error from the identification of spacelike and timelike $\widehat\Pi_{\rm had}(q^2)$ should be small except for the $b$ quark contribution, where the threshold is fairly close to the $q^2$ values of interest. Therefore we subtract the perturbative $b$ quark contribution evaluated for spacelike $q^2$ and add the corresponding value for timelike $q^2 = s$. We have checked that this ansatz is in excellent agreement with the numerical evaluation\footnote{We thank H.~Burkhardt for providing the numerical evaluation of (\ref{eqvacpolhad}) required for this comparison.} of (\ref{eqvacpolhad}) in the timelike region. Effects due to $\Upsilon$ resonances are described in detail below. The running $\alpha$ is then obtained from the real part of $\widehat\Pi$ through \begin{equation} \alpha(s) = \frac{\alpha}{1-{\rm Re}\widehat\Pi(s)}\,. \label{eqalpharunning} \end{equation} The relative shift in $\alpha(s)$ from the hadronic plus leptonic vacuum polarization is shown in Fig.~\ref{figalpha} as a function of $\sqrt{s}$. The size of the individual contributions is listed in Table~\ref{table2} for the reference energy $E_1=10.52$ GeV and for a few selected lower energies. The uncertainty in the cross section from our treatment of the hadronic vacuum polarization in the timelike region is estimated to be below two permille. \begin{table}[htb] \caption{Individual contributions to ${\rm Re}\widehat\Pi(s) \cdot 10^2$.} \begin{center} \begin{tabular}{|c||c|c|c|c|c|} \hline $\sqrt{s}/$GeV & $2m_{\tau}$ & 5 & 7 & 9 & 10.52 \\ \hline $e$ & 1.241 & 1.294 & 1.346 & 1.385 & 1.409 \\ $\mu$ & 0.415 & 0.468 & 0.520 & 0.559 & 0.583 \\ $\tau$ & -0.207 & -0.049 & 0.047 & 0.101 & 0.132 \\ had & 0.923 & 1.096 & 1.269 & 1.379 & 1.456 \\ \hline $1/\alpha(s)$ & 133.79 & 133.19 & 132.68 & 132.34 & 132.13 \\ \hline \end{tabular} \end{center} \label{table2} \end{table} \begin{figure}[htb] \begin{center} \leavevmode \epsfxsize=12.cm \epsffile[110 280 460 560]{alpha.ps} \vskip -3mm \caption[]{\label{figalpha} {\em The relative shift in $\alpha(s)$ as a function of $\sqrt{s}$ from the hadronic plus leptonic vacuum polarization as described in the text.}} \end{center} \end{figure} \vskip 0.5 cm \noindent {\bf (d) Narrow $\Upsilon$ resonances} \begin{table}[htb] \caption{Relative contributions from the $\Upsilon$ resonances to the leptonic cross section $\sigma(e^+ e^- \to \tau^+ \tau^-)$ for the energies $E_1 = 10.52$~GeV and $E_2 = 9.98$~GeV and two different values of the cutoff $m_{\rm min}$. Interference terms and radiative tails are listed separately. Also given are the continuum contributions and the resulting predictions for the total cross section.} \begin{center} \begin{tabular}{|c|c|c|c|c|c|c|} \hline &$\Upsilon(1)$&$\Upsilon(2)$&$\Upsilon(3)$& $\Upsilon(4)$&$\Upsilon(5)$&$\Upsilon(6)$\\ \hline\hline $M\ [GeV]$ & $9.460$ & $10.023$ & $10.355$ & $10.580$ & $10.865$ & $11.019$ \\ \hline $\Gamma_e\ [keV]$ & $1.32$ & $0.576$ & $0.476$ & $0.24$ & $0.31$ & $0.13$ \\ \hline $\Gamma_{\rm tot}\ [MeV]$ & $0.0525$ & $0.044$ & $0.0263$ & $23.8$ & $110$ & $79$ \\ \hline\hline \multicolumn{7}{|l|}{$E_1 = 10.52$ GeV, $m_{\rm min}=2m_{\tau}=3.554$ GeV:} \\ \hline Continuum: & \multicolumn{6}{c|}{$0.9256$ nb} \\ \hline Interf.: & $2.7\cdot 10^{-4}$ & $2.9\cdot 10^{-4}$ & $7.5\cdot 10^{-4}$ & $-1.04\cdot 10^{-3}$ & $-3.0\cdot 10^{-4}$ & $-9.7\cdot 10^{-5}$ \\ \hline $\sum$ interf.: & \multicolumn{6}{c|}{$-1.3\cdot 10^{-4}$} \\ \hline Rad.~tails~: & $5.5\cdot 10^{-4}$ & $2.5\cdot 10^{-4}$ & $8.1\cdot 10^{-4}$ & -- & -- & -- \\ \hline $\sum$ rad. tails: & \multicolumn{6}{c|}{$1.61 \cdot 10^{-3}$} \\ \hline Total: & \multicolumn{6}{c|}{$0.9269$ nb} \\ \hline\hline \multicolumn{7}{|l|}{$E_1 = 10.52$ GeV, $m_{\rm min}=5$ GeV:} \\ \hline Continuum: & \multicolumn{6}{c|}{$0.9033$ nb} \\ \hline Interf.: & $2.9\cdot 10^{-4}$ & $3.1\cdot 10^{-4}$ & $7.7\cdot 10^{-4}$ & $-1.07\cdot 10^{-3}$ & $-3.1\cdot 10^{-4}$ & $-9.8\cdot 10^{-5}$ \\ \hline $\sum$ interf.: & \multicolumn{6}{c|}{$-9.7\cdot 10^{-5}$} \\ \hline Rad.~tails~: & $5.7\cdot 10^{-4}$ & $2.6\cdot 10^{-4}$ & $8.3\cdot 10^{-4}$ & -- & -- & -- \\ \hline $\sum$ rad. tails: & \multicolumn{6}{c|}{$1.65 \cdot 10^{-3}$} \\ \hline Total: & \multicolumn{6}{c|}{$0.9047$ nb} \\ \hline\hline \multicolumn{7}{|l|}{$E_2 = 9.98$ GeV, $m_{\rm min}=2m_{\tau}=3.554$ GeV:} \\ \hline Continuum: & \multicolumn{6}{c|}{$1.0200$ nb} \\ \hline Interf.: & $6.4\cdot 10^{-4}$ & $-3.51\cdot 10^{-3}$ & $-4.5\cdot 10^{-4}$ & $-1.6\cdot 10^{-4}$ & $-1.6\cdot 10^{-4}$ & $-5.9\cdot 10^{-5}$ \\ \hline $\sum$ interf.: & \multicolumn{6}{c|}{$-3.70\cdot 10^{-3}$} \\ \hline Rad.~tails: & $1.08\cdot 10^{-3}$ & -- & -- & -- & -- & -- \\ \hline $\sum$ rad. tails: & \multicolumn{6}{c|}{$1.08\cdot 10^{-3}$} \\ \hline Total: & \multicolumn{6}{c|}{$1.0173$ nb} \\ \hline\hline \multicolumn{7}{|l|}{$E_2 = 9.98$ GeV, $m_{\rm min}=5$ GeV:} \\ \hline Continuum: & \multicolumn{6}{c|}{$0.9947$ nb} \\ \hline Interf.: & $6.8\cdot 10^{-4}$ & $-3.60\cdot 10^{-3}$ & $-4.6\cdot 10^{-4}$ & $-1.6\cdot 10^{-4}$ & $-1.6\cdot 10^{-4}$ & $-5.8\cdot 10^{-5}$ \\ \hline $\sum$ interf.: & \multicolumn{6}{c|}{$-3.75\cdot 10^{-3}$} \\ \hline Rad.~tails: & $1.11\cdot 10^{-3}$ & -- & -- & -- & -- & -- \\ \hline $\sum$ rad. tails: & \multicolumn{6}{c|}{$1.11\cdot 10^{-3}$} \\ \hline Total: & \multicolumn{6}{c|}{$0.9920$ nb} \\ \hline \end{tabular} \end{center} \label{table3} \end{table} \noindent Up to this point we have not included the contributions from the $\Upsilon$ resonances. As a consequence of their close proximity their effect will be enhanced and thus should be discussed separately. The Breit--Wigner amplitudes from the $\Upsilon$ resonances above and below the $B{\overline B}$ threshold interfere with the virtual photon amplitude and thus enhance or decrease the cross section by \begin{equation} \delta\sigma_{\rm int}= \frac{6\,\Gamma_e}{\alpha(s)\,s} \, \frac{M^3(s-M^2)}{(s-M^2)^2+\Gamma^2 M^2}\,\sigma_0\,. \label{eqbwint} \end{equation} The individual contributions are listed in Table~\ref{table3} for the two cms energies $E_1 = 10.52$~GeV and $E_2 = 9.98$~GeV after the convolution with initial state radiation. For $E_1$ these interference terms happen to cancel to a large extent, leaving a minute correction with a relative magnitude around $10^{-4}$. For $E_2$, however, due to the close proximity of $\Upsilon(2S)$ one receives a non--negligible negative contribution of about $-0.4$\%. This effect could become relevant in precision tests. \par Also the radiative tails of the resonances with $M < \sqrt{s}$ have to be taken into consideration. These can be easily calculated with the radiator function eq.~(\ref{eqGc}), taking either the full Breit--Wigner resonance \begin{equation} \delta\sigma_{\rm R} = \left(\frac{3\,\Gamma_e\,M}{\alpha(s)\,s}\right)^2 \frac{M^4}{(s-M^2)^2+\Gamma^2 M^2}\,\sigma_0 \label{eqbwtailfull} \end{equation} or the narrow width approximation \begin{equation} \delta\sigma_{\rm R} \Big|_{\rm NW} = \frac{9\,\Gamma_e^2}{\alpha(s)^2} \frac{M}{\Gamma} \,\pi\, \delta(s-M^2) \,\sigma_0\,. \label{eqbwtailnarroww} \end{equation} Both formulae lead to nearly identical results. The individual contributions are also listed in Table~\ref{table3}. In total they raise the lepton pair cross section by about 0.16\% or 0.11\% for $E_1$ and $E_2$, respectively. At this point a brief comment ought to be made concerning the treatment of $\Upsilon(4S)$. The approximation of neglecting the energy dependence of the width of the resonance is justified even for $E_1$, since the nominal width $\Gamma = 24$~MeV is significantly smaller than the difference between the cms energy $E_1$ and the mass of the resonance: $\Delta E = 60$~MeV. \pagebreak \noindent {\Large 3. The Total Cross Section for Hadron Production}\\ \noindent As stated already in the introduction the emphasis of this work is on the energy just below the bottom meson threshold. Virtual $b$ quark loops can be easily taken into account, and in fact it has been demonstrated that the hard mass expansion works surprisingly well for these diagrams --- not only far below but even close to the nominal $b$ quark threshold. On the other hand this energy is sufficiently far above the open charm threshold such that charm quark mass effects can be included through an expansion in powers of $m_c$, if quadratic and quartic terms are incorporated. With this method the production cross section can be predicted reliably not only in the high energy region but even relatively close to threshold through an expansion in $m^2/s$. This approach was suggested originally in \cite{quart,rhad}. The calculation of $R_{\rm had}$ to second order, including the full mass dependence \cite{hkt1} has demonstrated the nearly perfect agreement between approximate and exact result for the coefficient of the $\alpha_s^2$ term for $E_{\rm cm}$ above $4m$. In fact, even for $E_{\rm cm}$ around $3m$, which is around the lowest advisable value of the cutoff $m_{\rm min}=5$ GeV, the deviation of the $\alpha_s^2$ coefficient from the complete result leads to a difference of less than five percent in the rate. For the energy around 10 GeV this region contributes through the radiative tail only, and the approximation is thus adequate throughout. In total we thus use the following individual contributions \begin{equation} R = R_{\rm NS} + R_{\rm S} + \delta R_{m_b} + \delta R_{m_c} + \delta_{\rm QED}\,, \label{eqRhad} \end{equation} where \begin{eqnarray} R_{\rm NS} & = & \sum_{f = u,d,s,c} 3 \, Q_f^2 \left[ 1 + \frac{\alpha_s}{\pi} + 1.5245 \left(\frac{\alpha_s}{\pi}\right)^2 -11.52033 \left(\frac{\alpha_s}{\pi}\right)^3\, \right] \,,\nonumber\\ R_{\rm S} & = & -\left(\frac{\alpha_s}{\pi}\right)^3 \Big(\sum_{u,d,s,c} Q_f\Big)^2 \, 1.239 \ = \ -0.55091 \left(\frac{\alpha_s}{\pi}\right)^3 \,,\nonumber\\ \delta R_{m_b} & = & \sum_{f = u,d,s,c} 3 \, Q_f^2 \left(\frac{\alpha_s}{\pi}\right)^2 \frac{s}{{\overline m}_b^2} \left[ \frac{44}{675} + \frac{2}{135} \log \frac{{\overline m}_b^2}{s} \right] \,,\nonumber\\ \delta R_{m_c} & = & 3\,Q_c^2 \,12\,\frac{m_c^2}{s} \frac{\alpha_s}{\pi} \left[ 1 + 9.097 \frac{\alpha_s}{\pi} + 53.453 \left(\frac{\alpha_s}{\pi}\right)^2 \right] - 3 \sum_{f=u,d,s,c}Q_f^2\frac{m_c^2}{s} \left(\frac{\alpha_s}{\pi}\right)^3 6.476 \nonumber\\ & & + 3\,Q_c^2 \frac{m_c^4}{s^2} \left[ -6 -22 \frac{\alpha_s}{\pi} + \left( 141.329 - \frac{25}{6}\ln\frac{m_c^2}{s} \right) \left(\frac{\alpha_s}{\pi}\right)^2 \right] \nonumber\\ & &+3 \sum_{f=u,d,s,c} Q_f^2 \frac{m_c^4}{s^2} \left(\frac{\alpha_s}{\pi}\right)^2 \left[ -0.4749 - \ln\frac{m_c^2}{s} \right] -3\,Q_c^2 \frac{m_c^6}{s^3} \left[ 8 +\frac{16}{27} \frac{\alpha_s}{\pi} \left( 6\ln\frac{m_c^2}{s} + 155 \right) \right]\,, \nonumber\\ \delta_{\rm QED} & = & \sum_{f=u,d,s,c} 3\,Q_f^4 \frac{\alpha}{\pi} \frac{3}{4} \,. \nonumber \end{eqnarray} The formulae are evaluated for $n_f=4$ with $\alpha_s$ and the charm quark mass interpreted accordingly. For the massless case and the $m^2/s$ terms the results are available in third order, for the quartic and $m^6/s^3$ terms in second and first order $\alpha_s$, respectively. Tables which list the numerical values of the running quark masses and the magnitude of the individual contributions can be found in \cite{rhad,CKKRep} together with the details of the matching between the theories with $n_F = 4$ and $5$. \begin{figure}[htb] \begin{center} \leavevmode \epsfxsize=17.cm \epsffile[60 280 520 540]{rhad.ps} \vskip -3mm \caption[]{\label{fig3ab} {\em (a) $R_{\rm had}(s)$ as defined in eq.~(\ref{eqRhad}), and (b) only the contributions from the light ($u$, $d$ and $s$) quark currents for different values of $\alpha_s$.}} \end{center} \end{figure} The predictions for $R_{\rm had}$ as a function of $E_{\rm cm}$ are shown in Fig.~\ref{fig3ab} for three values of the strong coupling constant, $\alpha_s(M_Z^2) = 0.115$, $0.120$ and $0.125$. Fig.~\ref{fig3ab}a displays the contributions induced by the light plus charm quark currents (as defined in eq.~(\ref{eqRhad})), whereas in Fig.~\ref{fig3ab}b only the light ($u$, $d$ and $s$) quark current contributions are shown. Note, that this figure by definition does not contain QED--corrections from initial state radiation and vacuum polarization, but includes the tiny singlet terms, which cannot be attributed to one individual quark species and the final state QED--corrections $\delta_{\rm QED}$. The quark masses are chosen to be ${\overline m}_c({\overline m}_c) = 1.24$~GeV ($n_f=4$) and ${\overline m}_b({\overline m}_b) = 4.1$~GeV ($n_f=5$) corresponding to pole masses of $1.6$~GeV and $4.7$~GeV, respectively, if $\alpha_s^{(n_f=5)}(M_Z)=0.120$ is adopted. A variation of the charm quark mass around the default value by $+300$~MeV/$-300$~MeV changes $R_{\rm had}$ only by $+0.006/-0.004$ for $\sqrt{s} = 10.52$ GeV and by $+0.031/-0.028$ for $\sqrt{s} = 5$ GeV. Throughout this paper the QCD results are interpreted in fixed order $\alpha_s^3$ without any attempt to improve the formulae through the inclusion of guesses for higher order coefficients. An estimate of the scale dependence is easily obtained through the evaluation of a variant of eq.~(\ref{eqRhad}) where $R_{\rm NS}$ is calculated for a general t'Hooft scale $\mu^2$. Adopting $\alpha_s(10.5\ {\rm GeV}) = 0.177$ corresponding to $\alpha_s(M_Z) = 0.12$ and varying $\mu^2$ between $s/4$ and $4 s$ the predicted value of $R$ varies by -2 and +0.2 permille. This is well below the anticipated experimental precision. Alternatively we may include in eq.~(\ref{eqRhad}) an $\alpha_s^4$ term with the coefficient based on a recent estimate in \cite{Kataev}. This would lead to a decrease in $R$ by 1.7 permille, again far below the forseeable experimental accuracy. Let us now discuss the impact of initial state radiation, the running $\alpha$ and the $\Upsilon$ resonances on the hadronic cross section, as observed at around $10$~GeV under realistic experimental conditions. Lower energies contribute again through initial state radiation (eq.~(\ref{eqisrint})). If we restrict the cutoff $z_0$ to a value of $0.25$ which corresponds to a cut on the mass of the hadronic system of around $5$~GeV, eq.~(\ref{eqRhad}) for $R_{\rm had}$ can be applied also for $\sigma_0(z s')$ which appears in the integrand of (\ref{eqisrint}). At the same time this cutoff excludes the region of the broad charmonium resonances which have not been well explored up today. A cutoff around $z_0 = 0.7$ reduces the initial state radiation corrections further and eliminates contributions from the lower energy range completely. The vacuum polarization of the virtual photon has been discussed before for the case of the $\tau$ lepton production and is identical for the hadronic cross section. \begin{table}[htb] \caption{Relative contributions from the $\Upsilon$ resonances to the hadronic cross section $\sigma(e^+ e^- \to {\rm hadrons})$ for the energies $E_1 = 10.52$~GeV and $E_2 = 9.98$~GeV and the two different values of the cutoff $m_{\rm min} = 5$~GeV and 9~GeV. Interference terms and radiative tails are listed separately. Also given are the continuum contributions and the resulting predictions for the total cross section.} \begin{center} \begin{tabular}{|c|c|c|c|c|c|c|} \hline &$\Upsilon(1)$&$\Upsilon(2)$&$\Upsilon(3)$& $\Upsilon(4)$&$\Upsilon(5)$&$\Upsilon(6)$\\ \hline\hline \multicolumn{7}{|l|}{$E_1 = 10.52$ GeV, $m_{\rm min}=5$ GeV:} \\ \hline Continuum: & \multicolumn{6}{c|}{$3.2228$ nb} \\ \hline Interf.: & $2.9\cdot 10^{-4}$ & $3.1\cdot 10^{-4}$ & $7.7\cdot 10^{-4}$ & $-1.06\cdot 10^{-3}$ & $-3.1\cdot 10^{-4}$ & $-9.7\cdot 10^{-5}$ \\ \hline $\sum$ interf.: & \multicolumn{6}{c|}{$-1.0\cdot 10^{-4}$} \\ \hline Rad.~tails: & $5.88\cdot 10^{-3}$ & $5.38\cdot 10^{-3}$ & $1.212\cdot 10^{-2}$ & -- & -- & -- \\ \hline $\sum$ rad. tails: & \multicolumn{6}{c|}{$2.343\cdot 10^{-2}$} \\ \hline Total: & \multicolumn{6}{c|}{$3.2980$ nb} \\ \hline\hline \multicolumn{7}{|l|}{$E_1 = 10.52$ GeV, $m_{\rm min}=9$ GeV:} \\ \hline Continuum: & \multicolumn{6}{c|}{$2.8515$ nb} \\ \hline Interf.: & $4.0\cdot 10^{-4}$ & $3.7\cdot 10^{-4}$ & $8.9\cdot 10^{-4}$ & $-1.19\cdot 10^{-3}$ & $-3.3\cdot 10^{-4}$ & $-1.1\cdot 10^{-4}$ \\ \hline $\sum$ interf.: & \multicolumn{6}{c|}{$3.9\cdot 10^{-5}$} \\ \hline Rad.~tails: & $6.64\cdot 10^{-3}$ & $6.08\cdot 10^{-3}$ & $1.370\cdot 10^{-2}$ & $5\cdot 10^{-5}$ & -- & -- \\ \hline $\sum$ rad. tails: & \multicolumn{6}{c|}{$2.648\cdot 10^{-2}$} \\ \hline Total: & \multicolumn{6}{c|}{$2.9272$ nb} \\ \hline\hline \multicolumn{7}{|l|}{$E_2 = 9.98$ GeV, $m_{\rm min}=5$ GeV:} \\ \hline Continuum: & \multicolumn{6}{c|}{$3.5568$ nb} \\ \hline Interf.: & $6.7\cdot 10^{-4}$ & $-3.58\cdot 10^{-3}$ & $-4.6\cdot 10^{-4}$ & $-1.6\cdot 10^{-4}$ & $-1.6\cdot 10^{-4}$ & $-5.8\cdot 10^{-5}$ \\ \hline $\sum$ interf.: & \multicolumn{6}{c|}{$-3.74\cdot 10^{-3}$} \\ \hline Rad.~tails: & $1.145\cdot 10^{-2}$ & $9\cdot 10^{-5}$ & -- & -- & -- & -- \\ \hline $\sum$ rad. tails: & \multicolumn{6}{c|}{$1.154\cdot 10^{-2}$} \\ \hline Total: & \multicolumn{6}{c|}{$3.5845$ nb} \\ \hline\hline \multicolumn{7}{|l|}{$E_2 = 9.98$ GeV, $m_{\rm min}=9$ GeV:} \\ \hline Continuum: & \multicolumn{6}{c|}{$3.0666$ nb} \\ \hline Interf.: & $8.8\cdot 10^{-4}$ & $-4.1\cdot 10^{-3}$ & $-5.0\cdot 10^{-4}$ & $-1.7\cdot 10^{-4}$ & $-1.6\cdot 10^{-4}$ & $-6.1\cdot 10^{-5}$ \\ \hline $\sum$ interf.: & \multicolumn{6}{c|}{$-4.14\cdot 10^{-3}$} \\ \hline Rad.~tails: & $1.328\cdot 10^{-2}$ & $1.04\cdot 10^{-4}$ & -- & -- & -- & -- \\ \hline $\sum$ rad. tails: & \multicolumn{6}{c|}{$1.338\cdot 10^{-2}$} \\ \hline Total: & \multicolumn{6}{c|}{$3.0949$ nb} \\ \hline \end{tabular} \end{center} \label{table4} \end{table} \begin{figure}[htb] \begin{center} \leavevmode \epsfxsize=12.cm \epsffile[120 280 460 560]{haddiff.ps} \vskip -3mm \caption[]{\label{fig4} {\em Dependence of the hadronic cross section $\sigma(e^+ e^- \to {\rm hadrons})$ on the cutoff in the minimal invariant mass of the hadronic system, $m_{\rm min}$, for the two energies $E_1 = 10.52$ GeV, $E_2 = 9.98$ GeV and $\alpha_s(M_Z) = 0.120$.}} \end{center} \end{figure} The interference between the $\Upsilon$ resonances and the virtual photon is completely analogous to the leptonic case and leads to the same relative corrections. However, the contributions from the radiative tails of the Breit--Wigner amplitudes are distinctively different. Without initial state radiation the tail from one resonance is given by \begin{equation} \sigma_{\rm had, R} = \frac{12\,\pi\,\Gamma_e\,\Gamma_{\rm had}\,M^4}{s^3} \frac{M^2}{(s-M^2)^2+\Gamma^2 M^2}\,. \label{eqtailhad} \end{equation} For the $\Upsilon(4,\,5,\,6)$ resonances an energy dependent width $\Gamma(s)$ has to be used in (\ref{eqtailhad}), which varies rapidly in the region of interest. For $s$ below the $B{\overline B}$ threshold this leads to a drastic suppression of these contributions. In the narrow width approximation eq.~(\ref{eqtailhad}) implies \begin{equation} \sigma_{\rm had, R} \Big|_{\rm NW} = \frac{12\,\pi^2\,\Gamma_e\,\Gamma_{\rm had}}{s} \frac{M}{\Gamma}\,\delta(s-M^2) \,. \label{eqtailhadnw} \end{equation} The contributions of interference terms and radiative tails from the individual resonances to the hadronic cross section are listed in Table~\ref{table4} for the energies $E_1 = 10.52$ GeV and $E_2 = 9.98$ GeV and $\alpha_s(M_Z^2) = 0.120$. Also given are the continuum contribution and the sum. The cutoffs $m_{\rm min} = 5$~GeV and 9~GeV are adopted. These correspond to the minimal and maximal values recommended for a precision measurement of $R_{\rm had}$. The relative strength of the interference terms is identical for hadronic and leptonic final states. Again one observes the accidental cancellation at $10.52$~GeV and the dominance of the (negative) $\Upsilon(2)$ contribution at the energy $E_2 = 9.98$~GeV just below this resonance. Compared to the leptonic cross section (Table~\ref{table3}) the radiative tails are significantly more prominent --- a consequence of the relatively large direct, non--QED--mediated hadronic decay rates of the $\Upsilon$ resonances. This leads to a positive correction of 2.3\% and 1.2\% for $E_1$ and $E_2$, if $m_{\rm min} = 5$~GeV and to 2.7\% and 1.3\% if $m_{\rm min} = 9$~GeV, respectively. The difference in the strength of the $\Upsilon$ tails leads to an apparent variation of $R_{\rm had}$ around 1.3\%. The cutoff dependence of the cross section is illustrated in Fig.~\ref{fig4}. The strong dependence of the cross section for tight cuts is again clearly visible, suggesting a cut between about 5 and 9 GeV for $E_1$ and 5 and 8.5 GeV for $E_2$. \vskip 1 cm \noindent {\Large 4. Uncertainties}\\ \noindent From the comparison of different radiator functions for initial state radiation the theoretical uncertainty from this source can be estimated to be below one permille. The error induced through the present simplified treatment of the hadronic vacuum polarization in the timelike region is estimated around two permille and could easily be reduced even further, if required. The combined theoretical uncertainty from these and other effects is generously estimated below five permille. In addition photons from initial state radiation might be included in the invariant mass of the hadronic system or, conversely, photons from final state radiation may escape the detection. This ${\cal O}(\alpha)$ effect can only be evaluated for the concrete experimental analysis with the help of a Monte Carlo simulation. These theoretical uncertainties are significantly below the expected experimental error of roughly two percent, which is dominated by systematical uncertainties \cite{DaveBessonpriv}. \vskip 1 cm \noindent {\Large 5. Summary and Conclusions}\\ \noindent Precise predictions have been presented for the total cross section in the energy region explored presently by the CLEO experiment and at a future $B$ meson factory. The present sample of nearly one million hadronic events allows for a small statistical error. These measurements will determine the value for $\alpha_s$ under particularly clean conditions similar to the $Z$ line shape measurements but at a different energy. When compared to the experimental results from $Z$ decays, a determination of $R$ with a precision of 2.5\% would evidently demonstrate the running of $\alpha_s$ between 90 and 10 GeV. A precision of 0.3\% would be competitive with the $\alpha_s$ measurements from the $Z$ line shape which are based on the combined results of all four LEP experiments. \vskip 1 cm \noindent {\Large Acknowledgements}\\ \noindent We thank Fred Jegerlehner and Helmut Burkhardt for providing their programs and Fred Jegerlehner for important discussions concerning the hadronic vacuum polarization. The interest of Dave Besson in this study was essential for its completion. TT thanks the UK Particle Physics and Astronomy Research Council and the Royal Society for support. This work was supported by BMFT under Contract 057KA92P(0), and INTAS under Contract INTAS-93-0744. \def\app#1#2#3{{\it Act. Phys. Pol. }{\bf B #1} (#2) #3} \def\apa#1#2#3{{\it Act. Phys. Austr.}{\bf #1} (#2) #3} \defProc. LHC Workshop, CERN 90-10{Proc. LHC Workshop, CERN 90-10} \def\npb#1#2#3{{\it Nucl. Phys. }{\bf B #1} (#2) #3} \def\plb#1#2#3{{\it Phys. Lett. }{\bf B #1} (#2) #3} \def\prd#1#2#3{{\it Phys. Rev. }{\bf D #1} (#2) #3} \def\pR#1#2#3{{\it Phys. Rev. }{\bf #1} (#2) #3} \def\prl#1#2#3{{\it Phys. Rev. Lett. }{\bf #1} (#2) #3} \def\prc#1#2#3{{\it Phys. Reports }{\bf #1} (#2) #3} \def\cpc#1#2#3{{\it Comp. Phys. Commun. }{\bf #1} (#2) #3} \def\nim#1#2#3{{\it Nucl. Inst. Meth. }{\bf #1} (#2) #3} \def\pr#1#2#3{{\it Phys. Reports }{\bf #1} (#2) #3} \def\sovnp#1#2#3{{\it Sov. J. Nucl. Phys. }{\bf #1} (#2) #3} \def\jl#1#2#3{{\it JETP Lett. }{\bf #1} (#2) #3} \def\jet#1#2#3{{\it JETP Lett. }{\bf #1} (#2) #3} \def\zpc#1#2#3{{\it Z. Phys. }{\bf C #1} (#2) #3} \def\ptp#1#2#3{{\it Prog.~Theor.~Phys.~}{\bf #1} (#2) #3} \def\nca#1#2#3{{\it Nouvo~Cim.~}{\bf #1A} (#2) #3}
proofpile-arXiv_065-461
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{$D$-branes and the WZNW path integral.} The standard WZNW action on a group manifold $R$ reads \begin{equation} S(r)\equiv {1\over 4\pi}\int d\xi^+ d\xi^-\langle \partial_+ r~r^{-1}, \partial_- r~r^{-1}\rangle +{1\over 24\pi} \int d^{-1}\langle dr~r^{-1},[dr~r^{-1}, dr~r^{-1}]\rangle .\end{equation} Here $\xi^{\pm}$ are the standard lightcone variables on the world-sheet \begin{equation} \xi^{\pm}\equiv{1\over 2}(\tau \pm \sigma),\quad \partial_{\pm} \equiv \partial_{\tau}\pm\partial_{\sigma}\end{equation} and $\langle.,.\rangle$ denotes a non-degenerate invariant bilinear form on the Lie algebra ${\cal R}$ of $R$. The second term in the WZNW action is commonly referred to as the WZNW term and it provide the action with the antisymmetric tensor part. It is well-known that this antisymmetric tensor $B$ of the WZNW background is not globally defined (for compact groups) because the WZNW form $\Omega$ is a non-trivial cocycle in the third de Rham cohomology $H^3(R)$ of the group manifold $R$. Inspite of this, the classical WZNW theory is well defined for the case of closed strings. The reason is simple: Consider an evolving loop which sweeps out a cylindrical world-sheet $g(\sigma,\tau)$ on the group manifold. The variational problem requires fixing of the initial and final position of the loop and slightly varying the position of the cylinder between: $r(\sigma,\tau)\to r(\sigma,\tau)+\delta r(\sigma,\tau); ~\delta r\vert_{initial,final}=0$. The antisymmetric tensor part of the variation of the action can be thus written as \begin{equation} \int (r+\delta r)^*B-\int r^*B = \oint dB=\oint \Omega.\end{equation} The integral $\oint$ is taken over the volume interpolating between the world-sheets $r$ and $r+ \delta r$ and * means the pull-back of the map. We conclude that the variation of the action does indeed depend only on the WZNW three-from $\Omega$ and not on a choice of its potential $B$. Note that the interpolating volume is given unambiguously because the variation of the action is infinitesimal. A well known additional topological problem may occur if we wish to define a path integral for the WZNW theory of closed strings \cite{Wit}: Consider a set of fixed loops in $R$ and all world-sheets interpolating among them. We wish to evaluate the WZNW action $S$ of every world-sheet $s$, form an expression $\exp{iS}$ and sum up it over all interpolating world-sheets of arbitratry topology. Suppose we choose some reference interpolating world-sheet $s_{ref}$ and calculate its WZNW action $S_{ref}$ for some choice of the potential $B$. The action $S$ of any other world-sheet $s$ can be computed in the same way. It is tempting to conclude that the difference $S-S_{ref}$ does not depend on the choice of the potential $B$. Indeed, by using the same argument as in the variational problem, we easily see that the difference of the integral of $B$ over the both world-sheets is given solely in terms of the integral $\oint \Omega$ over the three-surface which interpolates between the world-sheets\footnote{We shall alway assume that the group manifold in question is simply connected. By Hurewicz isomorphism and the fact the second homotopy group of any Lie group vanishes we thus have that the second cohomology of the simply connected group manifold vanishes. This means that the interpolating three-surface always exists.}. But now the two world-sheets do not differ only infinitesimaly! It therefore seems that the interpolating three-surface is not given unambiguously. The way to get out of the trouble lies in comparing the quantity $S-S_{ref}$ for two non-homotopical three-surfaces interpolating between $s$ and $s_{ref}$. This difference is obviously given in terms of the integral $\oint \Omega$ over a three-cycle obtained by taking the difference of (or the sum of oppositely oriented) non-homotopical three-surfaces interpolating between $s$ and $s_{ref}$. Fortunately, the WZNW three-form $\Omega$ is an integer-valued cocycle \cite{Wit} in $H_3(R)$ hence it is enough to normalize action $S$ properly in order to ensure that the quantities $S-S_{ref}$ differ by a term $2\pi k, k\in Z$ for any two interpolating three-surfaces. These $2\pi k$ terms do not contribute to the path integral and, moreover, a dependence on the reference surface $s_{ref}$ results only in an unobservable change of the total phase of the path integral. We finish this little review by concluding that the WZNW path integral is well defined for the case of the interacting closed strings. Consider now a $D$-branes configuration in the group target $R$. By this we simply mean that there are two given submanifolds $D_i$ and $D_f$ of $R$ and open strings propagate on $R$ in such a way that their end-points $i$ and $f$ stick on the $D$-branes $D_i$ and $D_f$, respectively. We define the WZNW theory for this $D$-branes configuration by choosing two-forms $\alpha_i$ and $\alpha_f$, living respectively on $D_i$ and $D_f$ such that \begin{equation} d\alpha_{i(f)}= \Omega\vert_{D_i(D_f)} .\end{equation} In words: the exterior derivative of $\alpha_{i(f)}$ has to be equal to the restriction of the WZNW three form $\Omega$ to the $D$-brane $D_{i(f)}$. The construction of the $WZNW$ theory based on the triplet $(\Omega,\alpha_i,\alpha_f)$ goes as follows: Pick up an open string $r(\sigma,\tau)$ with the topology of an open strip. The variational problem requires fixing of the initial and the final positions of the string on the target. Consider now such a variation $\delta r(\sigma,\tau), \delta r(\sigma,\tau_{i,f})=0$. The both original open strip and its variation form together a closed strip (a `diadem'), whose edges lie on the opposite $D$-branes. We can define the variation $\delta S_{WZNW}$ of the WZNW term of the WZNW action by choosing an interpolating surface $\Sigma_{i(f)}\subset D_{i(f)}$ between the edges of the original and the varied strip. This variation then reads \begin{equation} \delta S_{WZNW}= \oint \Omega -\int_{\Sigma_i} \alpha_i - \int_{\Sigma_f} \alpha_f,\end{equation} where the $\oint \Omega$ is taken over the volume of the figure enclosed by $\Sigma_i$, $\Sigma_f$, the original strip and its variation. Note that this variation does not depend on the choice of the interpolating surface $\Sigma_{i(f)}$ because $d\alpha=\Omega\vert_D$ and all infinitesimal interpolating surfaces are mutually homotopic. Hence we conclude, that the classical WZNW theory of open strings with end-points on the $D$-branes is well defined in terms of the triplet $(\Omega,\alpha_i,\alpha_f)$. The reader may wish to have a more concrete idea of how to compute the WZNW action of a single strip. For a particular choice of the potential $B$ ($dB=\Omega$) the combination $\alpha-B$ on the $D$-brane is a closed form, hence, at least locally, it has a potential $A$ on $D$. The WZNW action $S$ for an open string configuration $r(\sigma,\tau)$ which sweeps out a two-surface $s$ in the target $R$ and respects the $D$-branes boundary conditions can now be written as follows \begin{equation} 4\pi S(r)=\int \langle \partial_+ r~r^{-1}, \partial_- r~r^{-1}\rangle + \int_s B +\int_{\delta s\cap D} A. \end{equation} Upon a change of \begin{equation} B\to B+d\lambda ,\end{equation} $A$ has to be replaced by \begin{equation} A-\lambda\vert_D . \end{equation} We may intepret the $A$-term of the action (6) as if there were equal and opposite charges on the end-points of the string which feel the electromagnetic fields $A_i$ and $A_f$ on the $D$-branes. This interpretation does not have an invariant meaning, however, because of the `gauge invariance' (7) and (8). Moreover it holds only locally. We stress that the {\it global} invariant description of the WZNW model for $D$-branes configuration is given in terms of the triplet $(\Omega,\alpha_i,\alpha_f)$. We remark that in general there is no natural {\it closed} two-form living on the $D$-branes. This is true only in the case if the restriction of the WZNW three-form $\Omega$ on the $D$-brane vanishes. Note also that if the $D$-brane is as many dimensional as the whole group target $R$ is, then the form $\alpha$ is nothing but some concrete choice of the potential $B$ which, however, may be different for the different end-points of the string. At the presence of the $D$-branes and open strings, the discussion of the string path integral is more involved as before. The group manifold will be always taken as simply connected and, for a while, we consider the case where also the $D$-branes are connected and simply connected. Now draw a general string diagram respecting the $D$-branes configurations. It is an interpolating world-sheet between a set of fixed open segments with end-points located on the $D$-branes and a fixed set of loops on the target $R$. Much as before, we can choose some reference interpolating world-sheet $s_{ref}$ and calculate its WZNW part of the action $S_{ref}$ for some choice of $B$ and $A$ according to the formula (6). Now we can take any other interpolating world-sheet $s$ and calculate its action $S$ in the same way. As in the case of the variational principle, the quantity $S-S_{ref}$ does not depend on the particular choice of $B$ and $A$ but only on the invariant globally defined triplet $(\Omega,\alpha_i,\alpha_f)$. The reason for this is the following: the union of the intersections $(\partial s_{ref}\cap D_{i(f)}) \cup (\partial s\cap D_{i(f)})$ is a contractible cycle in $D_{i(f)}$, hence it is a boundary of some two-surface $\Sigma_{i(f)}$. Now the union $s\cup s_{ref}\cup\Sigma_i\cup \Sigma_f$ is a two-boundary of some interpolating three-surface in the group manifold, because the second cohomology of the group manifold vanishes by assumption. Then the antisymmetric tensor (the WZNW term) part of $S-S_{ref}$ is defined by (5) where $\oint$ is taken over the interpolating three-surface. There occurs the same problem as for the closed strings, namely, the interpolating three-surfaces between $s$ and $s_{ref}$ do not have to be homotopically equivalent. This means that the quantity $S-S_{ref}$ may depend on the homotopy of the chosen interpolating three-surface. But if the ambiguity in $S-S_{ref}$ is only of the form $2\pi k, k\in Z$ then the term $\exp{i(S-S_{ref})}$ is unambiguous and the path integral is well defined. It is not difficult to find a cohomological formulation of the condition of the integer-valued ambiguity. All what we need is the notion of the relative singular homology $H_*(R,D_i\cup D_f)$ of the manifold $R$ with respect to its submanifolds $D_i$ and $D_f$ (with real coefficients). The relative chains are the elements of the vector space of the standard chains in $R$ factorized by its subspace of all chains lying in $D_i\cup D_f$. The operation of taking the boundary is the standard one. The corresponding homology is the relative singular homology $H_*(R,D_i\cup D_f)$. The triplet $(\Omega,\alpha_i,\alpha_f)$ can act on a relative cycle $\gamma$ by the following prescription \begin{equation} \langle (\Omega,\alpha_i,\alpha_f),\gamma\rangle\equiv \int_{\gamma}\Omega -\int_{D_i\cap\partial\gamma}\alpha_i -\int_{D_f\cap\partial\gamma}\alpha_f.\end{equation} If the cycle $\gamma$ is itself a boundary then the pairing vanishes because $\Omega$ is closed. Hence our triplet $(\Omega,\alpha_1,\alpha_f)$ is an element (cocycle) of the relative singular cohomology $H^*(R,D_i\cup D_f)$ because it vanishes on the boundary of any relative chain. Now we may conclude that if the cocycle $(\Omega,\alpha_1,\alpha_2)$ is integer-valued\footnote{The precise statement is as follows: The cocycle $(\Omega,\alpha_1,\alpha_2)$ is integer-valued, if it lies in the image of the natural map from the singular cohomology with integer coefficients to the singular cohomology with real coefficients.} the WZNW path integral is well-defined. Indeed, if we choose two non-homotopical three-surfaces interpolating between the world-sheets $s$ and $s_{ref}$ their oriented sum is a closed cycle in the relative singular homology and its pairing (9) with the triplet $(\Omega,\alpha_1,\alpha_2)$ is integer-valued. It turns out that we can extend our discussion to the case of connected but not necessarily simply connected $D$-branes. The main problem to be addressed is the fact that now the union of the intersections $(\partial s_{ref}\cap D_{i(f)}) \cup (\partial s\cap D_{i(f)})$ is not necessarily a contractible cycle in $D_{i(f)}$ (which means that $s\cup s_{ref}$ is a relative two-cycle but not a relative two-boundary). Thus the two-surface $\Sigma_{i(f)}$ does not have to exist and we cannot in general use the formula (5) in order to determine $\exp{i(S-S_{ref})}$. It may seem that we may take some reference world-sheet for each homotopy class of the one-chain $\partial s\cap D_{i(f)}$ and assign it an arbitrary reference phase. But there is still a consistency condition that under summing of the relative two-cycles (unions of $s$ and $s_{ref}$) the phases $\exp{iS}$ should be additive! Recall that we can unambiguously assign the $\exp{iS}$ to every relative two-boundary in such a way that this mapping is homomorphism $f$ from the group $B$ of relative two-boundaries (with integer coefficients) into the group $U$ of complex units (phases). The consistency condition means that there should exist an extension $\tilde f: Z\to U$ of this homomorphism defined on the group $Z$ of all relative two-cycles. We now prove that such an extension always exists because $U$ is the divisible group (this means that the equation $nx=a, a\in U, n\in {\bf N}$ has always a solution $x\in U$). Consider the group $H_f=Z+U/\{b-f(b),b\in B\}$. We have an exact sequence \begin{equation} 0\to U\to H_f \to H\to 0,\end{equation} where $H\equiv H_2(R, D_i\cup D_f)=Z/B$ and all homomorphisms are naturally defined. Suppose now that we do have an extension $\tilde f:Z\to U$ of the map $f:B\to U$. Such an extension enables us to write \begin{equation} H_f= H+U.\end{equation} In words: $H_f$ is a direct sum of $H$ and $U$. Indeed, for $z+c, z\in Z,c\in U$ we have \begin{equation} z+c = (z- \tilde f(z)) +(0 +c+\tilde f(z)).\end{equation} Evidently, the first term on the right hand side is from $Z$ and the second from $U$. The decomposition (12) is consistent with the factorization by $\{b-f(b), b\in B\}$ because $\tilde f$ is the homomorphism. The converse is also true: if we can write $H_f$ as the direct sum $H+U$ then there exists an extension $\tilde f:Z\to U$ which is a homomorphism. Indeed, consider $z\in Z$ and embed it naturally into $H_f$ i.e. $z\to z+0\in H_f$. $z+0$ can be decomposed as $y+g, y\in H, g\in U$ by assumption, hence we obtain a natural homomorphism from $Z$ into $U$: $z\to g$. This homomorphism is the extension of $f$ which we look for. Summarizing, if we prove that $H_f$ is the direct sum of $H$ and $U$, we are guaranteed that the extension $\tilde f:Z\to U$ always exists. But it is easy to prove this, by using the well-known result from the homological algebra that every extension of an (Abelian) group G by a divisible group $X$ is necessarily the direct sum of $G$ and $X$. In our case, we know from the exact sequence (10) that $H_f$ is the extension of $H$ by $U$. Therefore $H_f=H+U$, what was to be proved. \noindent {\it Notes}: \noindent 1. We have a certain freedom in writing $H_f$ as a direct sum of $H$ and $U$ which is described by the group of homomorphisms $Hom(H,U)$. The easiest way to see it is by noting that if we have an extension $\tilde f:Z\to U$ it can be modified by adding to it any homomorphism which vanishes on $b\in B$. Any such homomorphism is obviously from $Hom(H,U)$. The modified $\tilde f$ then gives another partition of $H_f$ into the direct sum of $H$ and $U$. \noindent 2. It may be instructive to relate the group $H$ of the relative two-cycles with the fundamental groups $\pi_1$ of the $D$-branes. We have a natural exact sequence \begin{equation} 0=H_2(R)\to H_2(R,D_i\cup D_f)\to H_1(D_i)+H_1( D_f)\to 0=H_1(R).\end{equation} Hence \begin{equation} H=H_1(D_i) +H_1(D_f)\end{equation} and \begin{equation} H_1(D_{i(f)})=\pi_1(D_{i(f)})/[\pi_1(D_{i(f)}),\pi_1(D_{i(f)})].\end{equation} The last equality is the Hurewicz isomorphism which holds due to the assumption that the $D$-branes are connected. \section{PL symmetries of WZNW models} For the description of the PL $T$-duality, we need the crucial concept of the Drinfeld double, which is simply a Lie group $D$ such that its Lie algebra ${\cal D}$ (viewed as a vector space) can be decomposed as the direct sum of two subalgebras, ${\cal G}$ and $\ti{\cal G}$, maximally isotropic with respect to a non-degenerate invariant bilinear form on ${\cal D}$ \cite{D}. It is often convenient to identify the dual linear space to ${\cal G}$ ($\ti{\cal G}$) with $\ti{\cal G}$ (${\cal G}$) via this bilinear form. From the space-time point of view, we have identified the targets of the mutually dual $\sigma$-models with the cosets $D/G$ and $D/\tilde G$ \cite{KS6}. Here $D$ denotes the Drinfeld double, and $G$ and $\tilde G$ two its mutually dual isotropic subgroups. In the special case when the decomposition $D=\tilde G G= G\tilde G$ holds globally, the corresponding cosets turn out to be the group manifolds $\tilde G$ and $G$, respectively \cite{KS2}. The actions of mutually dual $\sigma$-models are encoded in a choice of an $n$-dimensional linear subspace ${\cal R}$ of the $2n$-dimensional Lie algebra ${\cal D}$ of the double $D$ which is transversal to both ${\cal G}$ and $\ti{\cal G}$. The $\sigma$-model actions on the targets $D/G$ and $D/\tilde G$ have a similar structure; indeed, on $D/G$ we have \cite{KS6} \begin{equation} S={1\over 2}I(f)-{1\over 4\pi}\int d\xi^+ d\xi^- \langle \partial_+ f~f^{-1},R_-^a\rangle (M_-^{-1})_{ab}\langle f^{-1}\partial_- f,T^b\rangle,\end{equation} where $f\in D$ is some local section of the $D/G$ fibration which parametrizes the points of the coset. Recall \cite{KS6} that \begin{equation} M_{\pm}^{ab}\equiv \langle T^a ,f^{-1}R_{\pm}^b f\rangle\end{equation} and $R_-^a$ ($R_+^a$) are vectors of an orthonormal basis of ${\cal R}$ (${\cal R}^{\perp}$): \begin{equation} \langle R_{\pm}^a,R_{\pm}^b\rangle= \pm\delta^{ab},\qquad \langle R_+^a,R_-^b\rangle=0.\end{equation} The action of the dual $\sigma$-model on the coset $D/\tilde G$ has the same form; just the generators $T^a$ of ${\cal G}$ are replaced by the generators $\tilde T_a$ of $\ti{\cal G}$ and $f$ will parametrize $D/\tilde G$ instead of $D/G$. We have referred to the $\sigma$-models of the form (16) as those having a PL symmetry \cite{KS6}. There is an important feature of such models, namely, their field equations can be written as the zero curvature condition valued in the algebra ${\cal G}$. Indeed, \begin{equation} d\lambda-\lambda^2=0,\end{equation} where \begin{equation} \lambda =\lambda_+ d\xi^+ +\lambda_- d\xi^-\end{equation} and \begin{equation} \lambda_{\pm}=-\langle \partial_{\pm} f~f^{-1},R_{\mp}^a\rangle (M_{\mp}^{-1})_{ab}T^b.\end{equation} So far we have been reviewing the results of \cite{KS6}; now a new observation comes: If the subspace ${\cal R}$ is itself a Lie algebra of a compact subgroup $R$ of the double $D$ then the model (16) is essentially the WZNW model on the target $R$ for the both choices $D/G$ and $D/\tilde G$! The argument goes in two steps: \noindent 1. ${\cal R}$ can be transported by the right action to the tangent space of every point of the double. Because ${\cal R}$ is the subalgebra, the distribution of the planes ${\cal R}$ in the tangent bundle of the double is integrable and it foliates the double into fibration with fibres $R$ and basis $R\backslash D$. Since ${\cal R}$ is transversal to the both ${\cal G}$ and $\ti{\cal G}$ (which means that it intersects ${\cal G}$ and $\ti{\cal G}$ only in $O$) , any fiber of the $R$ fibration either intersects the fiber $G$ (or $\tilde G$) in some finite subgroup $R\cap G$ of $R$ or does not intersect it at all. The latter cannot be true, however, if the group $R$ is compact. Indeed, $R$ acts on $D/G$ by the left action. The $R$ orbit of the element of $D/G$ which has the unit element of $D$ on its fiber is open. Since $R$ is compact this orbit must be also closed which for connected doubles imply that this orbit is the whole $D/G$. In other words, there always exists an intersection of $R$ and $G$. The argument for $D/\tilde G$ is the same. If the finite subgroups $R\cap G$ and $R\cap \tilde G$ have only one element for both fibers $G$ and $\tilde G$, respectively, it si not dificult to see that the both cosets $D/G$ and $D/\tilde G$ can be globally identified with $R$. In general, the cosets $D/G$ and $D/\tilde G$ can be identified with the discrete cosets $R/R\cap G$ and $R/R\cap \tilde G$, respectively. \noindent 2. For simplicity, consider only the case when $R$ can be directly identified with $D/G$ and $D/\tilde G$. In this case, we can choose the field $f(\sigma,\tau)$ in (16) to have values in $R$. Note that we can choose the basis $R_-^a$ dependent on $f$ in such a way that the combinations $f^{-1}R_-^a f$ are $f$ independent. Then we can choose the basis $T^a$ in such a way that $M_-(f)$ is the identity matrix. We have \begin{equation} \langle \partial_+ f ~f^{-1}, R_-^a\rangle =\langle f^{-1} \partial_+ f, f^{-1}R_-^a f\rangle\equiv (f^{-1} \partial_+ f)^a\end{equation} and \begin{equation} \langle f^{-1}\partial_- f , T^a\rangle= \langle f^{-1}\partial_- f,f^{-1}R_-^c f\rangle M_-^{ca}=(f^{-1}\partial_- f)^a,\end{equation} because $M_-$ is the identity matrix. Putting (16),(22) and (23) together, we obtain \begin{equation} S={1\over 2}I(f)-{1\over 4\pi}\int d\xi^+ d\xi^- (\partial_+ f ~f^{-1})^a \delta_{ab}(\partial_- f ~f^{-1})^b=-{1\over 2} I(f^{-1} ).\end{equation} We conclude, that the mutually dual $\sigma$-models on the cosets $D/G$ and $D/\tilde G$ are the same, being equal to the WZNW model on $R$. In general, $D/G$ ($D/\tilde G$) model is WZNW model on the target $R/R\cap G$ ($R/R\cap \tilde G$). \noindent {\it Notes}: \noindent 1. The fact that the both models $D/G$ and $D/\tilde G$ may be identical does not mean at all that the duality transformation is trivial. In fact, the PL $T$-duality always implies an existence of a non-trivial non-local transformation on the phase space of the $WZNW$ model. We shall explicitly describe this transformation in the next section. \noindent 2. It often happens (cf. section 4) that a compact group $R$ can be embedded in many inequivalent ways into various Drinfeld doubles in such a way that the both cosets $D/G$ and $D/\tilde G$ can be identified with $R$. In this case we have the abundance of the Poisson-Lie symmetries of the same WZNW model on the group manifold $R$, each of them corresponding to the double into which $R$ is embedded. \section{$D$-branes in WZNW models} \subsection{General discussion} For the further discussion of the $D$-branes, it is convenient to recall \cite{KS6} the common `roof' of the both models described by (16). They can be derived form the first order Hamiltonian action for field configurations $l(\sigma,\tau)\in D$: $$ S[l(\tau,\sigma)]= $$ \begin{equation} ={1\over 8\pi}\int \biggl\{\langle \partial_{\sigma} l~l^{-1},\partial_{\tau} l~l^{-1}\rangle+ {1\over 6}d^{-1}\langle dl~l^{-1},[dl~l^{-1}, dl~l^{-1}]\rangle -\langle \partial_{\sigma} l l^{-1},A\partial_{\sigma} l l^{-1}\rangle \biggl\}.\end{equation} Here $A$ is a linear idempotent self-adjoint map from the Lie algebra ${\cal D}$ of the double into itself. It has two equally degenerated eigenvalues $+1$ and $-1$, and the corresponding eigenspaces are just ${\cal R}^{\perp}$ and ${\cal R}$ respectively. As it stands, the action (25) is well defined only for the periodic functions of $\sigma$ because of the WZNW term. This restriction corresponds to the case of closed strings \cite{KS6} . The $\sigma$-model actions (16) are obtained from the duality invariant first order action (25) as follows: Consider the right coset $D/G$ and parametrize it by the elements $f$ of $D$ \footnote{If there exists no global section of this fibration, we can choose several local sections covering the whole base space $D/G$.}. With this parametrization of $D/G$ we may parametrize the surface $l(\tau,\sigma)$ in the double as follows \begin{equation} l(\tau,\sigma)= f(\tau,\sigma)g(\tau,\sigma),\quad g\in G.\end{equation} The action $S$ then becomes $$ S(f,\Lambda\equiv \partial_{\sigma} g g^{-1})={1\over 2}I(f) -{1\over 2\pi} \int d\xi^+ d\xi^- \biggl\{\big\langle \Lambda - {1\over 2} f^{-1}\cdot- f , \Lambda -{1\over 2}f^{-1}\cdot- f \big\rangle$$ \begin{equation} +\langle f\Lambda f^{-1} +\partial_{\sigma} f f^{-1}, R_-^a\rangle\langle R_-^a , f\Lambda f^{-1} +\partial_{\sigma} f f^{-1}\rangle\biggl\}.\end{equation} Now it is easy to eliminate $\Lambda$ from the action (27) and finish with the $\sigma$-model action (16). In the case of the coset $D/\tilde G$, the procedure is exactly analoguous. Consider the case of open strings for a generic double $D$ with vanishing second cohomology. In our previous paper on the subject \cite{KS4}, we have studied only the perfect doubles (cf. footnote 3) nevertheless we can easily generalize the construction. Let $F$ be a simply connected subgroup of the double $D$ whose Lie algebra ${\cal F}$ is isotropic with respect to the bilinear form on ${\cal D}$. This subgroup, as a manifold, can be shifted by the right action of some element $d\in D$ (note that all non-equivalent shifts are parametrized by the coset $F\backslash D$). We declare that the manifolds $F\hookrightarrow D$ and $Fd\hookrightarrow D$ are $D$-branes in the double $D$. Consider now oriented open strings in $D$ with the initial end-points on $F$ and the final end-points on $Fd$. Their dynamics in the bulk is governed by the action (25) which contains the WZNW term. As we have learnt in the previous section such an action is well-defined provided we choose some two-forms on the $D$-branes such that the exterior derivative of them is equal to the restriction of the $WZNW$ three-form on the $D$-branes. In our present case, this restriction of the WZNW form vanishes in either of our $D$-branes because $F$ and $Fd$ are the isotropic surfaces in $D$. Thus we have to choose some closed two forms on $F$ and $Fd$; we choose them to vanish identically. We summarize that our open string dynamics is fully defined by the action (25), the $D$-branes boundary conditions and the vanishing two-forms on the $D$-branes. Much as in the closed string case, we can derive the open string $\sigma$-model dynamics on the cosets $D/G$ and $D/\tilde G$ from (25) and the $D$-branes data on the double; for concreteness let us consider the coset $D/G$: As we have learnt in section 2, the WZNW model for open strings is fully defined if we manage to compute the WZNW action of the `diadem'. Recall that the diadem is composed of two evolving open string world-sheets which are glued together at some initial and final times. The edges of the diadem , swept by the end-points of the open strings, lie in their corresponding $D$-branes. Consider now the diadem in the double. We can choose some two-surface $\Sigma$ ($\Sigma_d$) in the $D$-brane $F$ ($Fd$) whose boundary is just the edge of the diadem lying in $F$ ($Fd$). The diadem together with the surfaces $\Sigma$ and $\Sigma_d$ form a boundary of some three-dimensional domain $\gamma$. We may write the action $S$ of the model (25) as \begin{equation} S=S_0+S_{WZNW},\end{equation} where $S_{WZNW}$ contains solely the term with the WZNW three-form $c$ on $D$. Hence, the action of the diadem can be written as\footnote{Note that we have included the factor $1/6$ from (25) in the definition of $c$.} \begin{equation} S=S_0 + {1\over 8\pi}\int_{\gamma}c.\end{equation} Again, consider the parametrization of $D/G$ by the elements $f$ of $D$. A surface $l(\tau,\sigma)$ in the double (respecting the $D$-branes boundary conditions), can be written as \begin{equation} l(\tau,\sigma)= f(\tau,\sigma)g(\tau,\sigma),\quad g\in G.\end{equation} The decomposition (30) induces two maps from $D$ into $D$: $f(l)=f$ and $g(l)=g$. Consider now the Polyakov-Wiegmann (PW) formula \cite{PW} \begin{equation} (fg)^*c=f^*c +g^*c -d\langle f^*(l^{-1}dl) \stackrel {\wedge}{,}g^*(dl ~l^{-1}) \rangle,\end{equation} where, as usual, $*$ denotes the pull-back of the forms under the mappings to the group manifold $D$. By using the PW formula, we can rewrite (29) as \begin{equation} S=S_0(fg) +{1\over 8\pi}\int_{\gamma} f^*c -{1\over 8\pi}\int_{diad\cup \Sigma \cup \Sigma_d}\langle f^*(l^{-1}dl)\stackrel {\wedge}{,} g^*(dl~l^{-1})\rangle.\end{equation} Note that $g^*c$ vanishes because of the isotropy of $G$. The action $S$ now becomes $$ S(f,\Lambda\equiv \partial_{\sigma} g g^{-1})={1\over 2\pi}\int_{diad}\biggl\{ {1\over 4}\langle \partial_+ f~f^{-1},\partial_- f~f^{-1}\rangle$$ $$ -\big\langle \Lambda - {1\over 2} f^{-1}\cdot- f , \Lambda -{1\over 2}f^{-1}\cdot- f \big\rangle +\langle f\Lambda f^{-1} +\partial_{\sigma} f f^{-1}, R_-^a\rangle\langle R_-^a , f\Lambda f^{-1} +\partial_{\sigma} f f^{-1}\rangle\biggl\}$$ \begin{equation} +{1\over 8\pi}\int_{\gamma}f^*c -{1\over 8\pi}\int_{\Sigma\cup \Sigma_d}\langle f^*(l^{-1} dl)\stackrel {\wedge}{,} g^*(dl~l^{-1} )\rangle.\end{equation} Of course, this is a similar expression as before (cf. (27)). However, the field $f$ respects different boundary conditions. A configuration $f$ is an open string configuration; its end-points stick on $D$-branes $D_i$ and $D_f$ in $D/G$ which are obviously obtained just by projecting the $D$-branes $F$ and $Fd$ from the double D into the basis $D/G$ parametrized by the section $f$. Now we have to realize that upon varying $\Lambda$ the last term in (33) vanishes! This follows from the isotropy of $F$, $Fd$ and $G$. Indeed, if we have $fg\in F$ ($fg\in Fd$) and vary $g\to g\delta g$ at fixed $f$ in such a way that $fg\delta g\in F$ ($fg\delta g \in Fd$), we observe that the last term in (33) does not change\footnote{It is easy to see that $\delta g\in F\cap G$ ($\delta g\in F\cap dGd^{-1}$).}. Hence we can eliminate the field $\Lambda$ from (33) in the same way as from (27). The result is $$ S={1\over 8\pi}\int_{diad} \biggl\{\langle \partial_+ f~f^{-1},\partial_- f~f^{-1}\rangle -2 \langle \partial_+ f~f^{-1},R_-^a\rangle (M_-^{-1})_{ab}\langle f^{-1}\partial_- f, T^b\rangle\biggl\}$$ $$ +{1\over 8\pi}\int_{\gamma}f^*c -{1\over 8\pi}\int_{\Sigma\cup \Sigma_d}\langle f^*(l^{-1}dl)\stackrel {\wedge}{,} g^*(dl~l^{-1}) \rangle.$$ Consider again the special situation in which the subspace ${\cal R}\equiv {\rm Span} R_-$ is the Lie algebra of the compact group $R$, moreover, $R$ can be directly identified with $D/G$ and $D/\tilde G$. Recall, that upon transporting ${\cal R}$ by the right action everywhere onto the double, we get the fibration of $D$ with the fibers $R$ and the basis $R\backslash D$. With some abuse of the notation, the fiber crossing the unit element of the double we shall also denote as $R$. We choose the parametrization of the double as follows \begin{equation} l=rg,\quad r\in R,\quad g\in G.\end{equation} This parametrization holds for every element $l$ of the double and is unique by the assumption. Note that the restriction of the WZNW three-form $c$ gives just the WZNW three-form $c_R$ on $R$. It is easy to see that the $D$-branes $D_i$ and $D_f$ in $R$, being the projections of $F$ and $Fd$ to $R$, can be identified with the cosets $F/F\cap G$ and $F/F\cap dGd^{-1}$ respectively. On the other hand we have just seen (cf. footnote 9) that the variation $\delta g\in F\cap G$ ($\delta g \in F\cap dGd^{-1}$) leaves intact the two-form $\omega \equiv (1/8\pi)\langle r^*(l^{-1}dl)\stackrel {\wedge}{,} g^*(dl~l^{-1} )\rangle$ on $F$ ($Fd$). This means that this two-form is a pull-back of some two-form $\alpha_i$ ($\alpha_f$) from the $D$-brane $D_i$ ($D_f$). Of course, the notation is not accidental; the two-forms $\alpha_{i(f)}$ are precisely those appearing in (5). It is not difficult to find an explicit expression for $\alpha_{i(f)}$. For this, consider a map $k_i$ ($k_f$) from $D_i$ ($D_f$) into $G$ such that \begin{equation} rk_{i(f)}(r) \in F(Fd), \quad r\in D_{i(f)}.\end{equation} In general, the mapping $k_i$ ($k_f$) is not defined unambiguously but it locally always exists since $D_i$ ($D_f$) is just the projection of $F$ ($Fd$) on $R$. Because two-form $\omega$ on $F$ ($Fd$) is invariant under the variations from $F\cap G$ ($F\cap dGd^{-1}$) we can locally\footnote{The two-form $\alpha_{i(f)}$ is defined {\it globally} on $D_{i(f)}$ only the explicit expression for it in terms of $k_{i(f)}$ may, in general, be written only locally.} write \begin{equation} \langle r^*(l^{-1}dl)\stackrel {\wedge}{,} g^*(dl~l^{-1})\rangle\vert_{F(Fd)}= r^*\langle dr ~r^{-1}\stackrel {\wedge}{,} k_{i(f)}(r)^{-1}dk_{i(f)}(r)\rangle.\end{equation} In other words, (36) is true independently of the choice of the map $k_{i(f)}$. Thus in our special situation, the action of the diadem can be written as $$ S=-{1\over 8\pi}\int_{diad} \langle \partial_+ r~r^{-1},\partial_- r~r^{-1}\rangle + {1\over 8\pi}\int_{r(\gamma)}c_R $$ \begin{equation} -{1\over 8\pi}\int_{D_i}\langle dr ~r^{-1} \stackrel {\wedge}{,} k_i(r)^{-1}dk_i(r)\rangle -{1\over 8\pi}\int_{D_f}\langle dr ~r^{-1} \stackrel {\wedge}{,} k_f(r)^{-1}dk_i(r)\rangle.\end{equation} We can read $\alpha_{i(f)}$ off directly from (37): \begin{equation} \alpha_{i(f)}= {1\over 8\pi}\langle dr ~r^{-1} \stackrel {\wedge}{,} k_{i(f)}(r)^{-1}dk_{i(f)}(r)\rangle.\end{equation} It remains to prove that \begin{equation} d\alpha_{i(f)}={1\over 8\pi} c_R\vert_{D_{i(f)}}.\end{equation} This is easy: take the PW formula (31) and restrict all forms in it on the $D$-brane $F$ ($Fd$) in the double. Then the form $c$ vanishes by the isotropy of $F$ ($Fd$). Hence \begin{equation} r^*c\vert_{F(Fd)}=d\langle r^*(l^{-1}dl)\stackrel {\wedge}{,} g^*(dl~l^{-1}) \rangle\vert_{F(Fd)}=r^*d\langle dr ~r^{-1}\stackrel {\wedge}{,} k_{i(f)}(r)^{-1}dk_{i(f)}(r)\rangle,\end{equation} where the last equality follows from (36). Thus, upon removing the pull-back map $r^*$, we conclude that \begin{equation} {1\over 8\pi}c_R\vert_{D_{i(f)}}={1\over 8\pi}d\langle r^{-1} dr \stackrel {\wedge}{,} dk_{i(f)}(r)k_{i(f)}(r)^{-1}\rangle= d\alpha_{i(f)}.\end{equation} \noindent {\it Remarks}: \noindent 1. The model (37) has the `wrong' sign in front of its first term. Upon the change of variables $r\to r^{-1}$ it gives the standard WZNW model on the group manifold $R$ (cf. (1)). The $D$-branes $D_{i(f)}$ and the two forms $\alpha_{i(f)}$ on them have to be transformed correspondingly. \noindent 2. The geometry of the dual $D$-branes in $D/\tilde G$ is obtained in the same way as in the case $D/G$; it is enough to replace everywhere $G$ by $\tilde G$. \noindent 3. We should mention that the Kiritsis-Obers duality \cite{KO} fits in our formalism. The double is the direct product of a compact group $R$ with itself and the invariant bilinear form in the direct sum of the Lie algebras ${\cal R}+{\cal R}$ is the difference between the Killing-Cartan forms on each algebra. Hence, the diagonal embedding of $R$ in $R\times R$ is isotropic. So it is the embedding in which second copy of ${\cal R}$ is twisted by some outer automorphism. The resulting duality is a $D$-branes $D$-branes duality, i.e. the $D$-branes have never the dimension of the group manifold. \subsection{The classical solvability} We wish to find the complete solution of the field equations of the model (25) submitted to the $D$-branes boundary conditions. It is not difficult to do that. The bulk equations following from (25) read \begin{equation} \langle \partial_{\pm}l ~l^{-1}, {\cal R}_{\mp}\rangle=0.\end{equation} We already know that after integrating away $g$ from the decomposition (34) we get the WZNW model on $R$, hence, the solution $l$ of (25) must look like \begin{equation} l(\sigma,\tau)=r_-(\xi^-)r_+(\xi^+)g(\xi^+).\end{equation} The first two multiplicative terms on the right-hand-side follow from the known bulk solution of the WZNW model on $R$ and the fact that $g$ is only a function of $\xi^+$ follows from Eqs. (21). Putting \begin{equation} h(\xi^+)\equiv r_+(\xi^+)g(\xi^+)\end{equation} and inserting $l=r_-(\xi^-)h(\xi^+)$ into (37), we obtain \begin{equation} \partial_+h ~h^{-1}\in {\cal R}_+\equiv {\cal R}^{\perp}.\end{equation} Here we have used the fact that ${\cal R}_-(\equiv {\cal R})$ is the Lie algebra of $R$. We conclude that every bulk solution of (25) look like \begin{equation} l=r_-(\xi^-)h(\xi^+), \qquad \partial_+h~h^{-1}\in {\cal R}_+.\end{equation} It is important to note that ${\cal R}_+\equiv{\cal R}^{\perp}$ does not have to be a Lie subalgebra of ${\cal D}$; in general it is just a linear subspace of ${\cal D}$. Now we can take into account the effect of the boundary conditions. Recall that the initial point of the open string ($\sigma=0$) should stick on the $D$-brane $F$ in the double and the final point ($\sigma=\pi$) on the $D$-brane $Fd$; $d$ is a fixed element of the double $D$. These two conditions can be rewritten as follows \begin{equation} r_-(\tau)h(\tau)=f_i(\tau),\qquad r_-(\tau-\pi)h(\tau)=f_f(\tau)d,\end{equation} where $f_i$ and $f_f$ are some functions with values in the group $F$. It follows that \begin{equation} h^{-1}(\tau-\pi)h(\tau)=f_i^{-1}(\tau-\pi)f_f(\tau)d.\end{equation} By differentiating Eq. (48) with respect to $\tau$ we obtain $$ -dh(\tau-\pi)h^{-1}(\tau-\pi)+dh(\tau)h^{-1}(\tau)=$$ \begin{equation} =h(\tau-\pi)[-f_i^{-1}(\tau-\pi)df_i(\tau-\pi) +f_i^{-1}(\tau-\pi) df_f(\tau)f_f^{-1}(\tau)f_i(\tau-\pi)]h^{-1}(\tau-\pi).\end{equation} Now we can bracket (49) with ${\cal R}$ which gives \begin{equation} df_f(\tau)f_f^{-1}(\tau)-df_i(\tau-\pi)f_i^{-1}(\tau-\pi)=0.\end{equation} For deriving (50), we have used Eq. (47) and the fact that the Lie algebra ${\cal F}$ of $F$ is transversal to ${\cal R}^{\perp}$. By inserting (50) back in (49) we get a very important relation \begin{equation} dh(\tau+\pi)h^{-1}(\tau+\pi)=dh(\tau)h^{-1}(\tau).\end{equation} It expresses the periodicity of the ${\cal R}^{\perp}$-valued `connection' $dh~h^{-1}$. The monodromy of this `connection is also constrained; indeed, from (50) and (48) we conclude that \begin{equation} h^{-1}(\tau-\pi)h(\tau)=fd,\end{equation} where $f$ is some constant element of $F$. In words: the monodromy $h^{-1}(\tau-\pi)h(\tau)$ is an element of the double $D$ which is equivalent to $d$ in the sense of the coset $F\backslash D$. \noindent {\it Summary}: The space of the solutions of the field equations (42) submitted to the $D$-branes boundary conditions (47) is given by an arbitrary element $p$ of the double $D$ and a periodic field $\rho(\xi^+)(\equiv dh~h^{-1}(\xi^+))$ with values in the subspace ${\cal R}^{\perp}$ of ${\cal D}$ and with the monodromy \begin{equation} P\exp{\int_{\tau -\pi}^{\tau}d\tau' \rho(\tau')}\equiv h^{-1}(\tau-\pi)h(\tau)\end{equation} equivalent to $d$ in the sense of the coset $F\backslash D$. Of course, $P$ in (53) means the ordered exponent. The full solution $l(\sigma,\tau)$ is then reconstructed as follows: take $\rho(\xi^+)$ and $p\in D$ and construct \begin{equation} h(\xi^+)=P\exp{\{\int_{\xi_0^+}^{\xi^+}d\xi^{+'} \rho(\xi^{+'})\}}\times p.\end{equation} Obviously, the choice of $\xi^+_0$ is irrelevant and can be compensated by the corresponding change of $p$. Now we can reconstruct $r_-(\xi^-)$ by decomposing $h(\xi^-)$ as \begin{equation} h(\xi^-)=r_-^{-1}(\xi^-)f(\xi^-), \qquad r\in R,\quad f\in F. \end{equation} This decomposition is unique, because $R$ can be globally identified with $D/F$. Finally \begin{equation} l(\xi^+,\xi^-)=r_-(\xi^-)h(\xi^+).\end{equation} It remains to recover from the solution (56) on the double the solutions of the $\sigma$-models on the cosets $D/G$ and $D/\tilde G$. Recall that the both $\sigma$-models are the WZNW models on the group manifold $R$. In the $D/G$ case we have to decompose $h$ as \begin{equation} h(\xi^+)=r_+(\xi^+)g(\xi^+), \qquad r\in R, \quad g\in G,\end{equation} while in the $D/\tilde G$ case as \begin{equation} h(\xi^+)=\tilde r_+(\xi^+)\tilde g(\xi^+), \qquad \tilde r\in R, \quad \tilde g\in \tilde G.\end{equation} Because $r_+\neq \tilde r_+$ we indeed obtain a nontrivial map from the phase space of the WZNW model with one set of the $D$-branes boundary conditions into the phase space of the same WZNW model but with the dual $D$-branes boundary conditions. The both phase spaces can be identified with the set of all solutions $l(\xi^+,\xi^-)$ on the double. The system (25) is already written in the Hamiltonian form, hence the mapping between the phase spaces is a canonical transformation. \noindent {\it Note}: It is interesting that the both left movers $r_+(\xi^+)$ and right movers $r_-(\xi^-)$ are obtained from the master function $h(\xi^+)$ in a very similar way. Recall that \begin{equation} h(\xi^+)= r_+(\xi^+) g(\xi^+), \qquad h(\xi^-)= r_-^{-1}(\xi^-) f(\xi^-), \quad g\in G,~f\in F.\end{equation} In particular, if $G=F$ then the left and the right movers of the $R$ WZNW model are given by the same function , i.e. \begin{equation} r(\sigma,\tau)=r_-(\xi^-)r_-^{-1}(\xi^+).\end{equation} This means that the initial point $\sigma=0$ of the string sits at the origin of the group $R$ for all times. Indeed, the corresponding $D$-brane is just the group origin, being the projection of $F=G$ along $G$. \subsection{Interacting $D$-brane diagrams} Given a $D$-brane configuration on the target $R$ we can in principle compute the WZNW path integral over all topologically non-trivial world-sheets interpolating between a set of fixed open string segments with end-points sitting on the $D$-branes and a set of fixed loops in the target $R$. We postpone the evaluation of some of such diagrams (like the open string propagator) to a forth-coming publication, here we just discuss whether there are some topological obstructions in doing that possibly coming from the WZNW term $\Omega$ and the two-forms $\alpha_i$ and $\alpha_f$ defined on the $D_i$ and $D_f$ by (38). We have learned in the section 2 that the WZNW path integral is well-defined if the triplet $(\Omega,\alpha_i,\alpha_f)$ is an integer-valued cocycle in the relative singular cohomology of the group manifold $R$ with respect to its submanifold $D_i\cup D_f$. In general, we have found it to be a difficult topological problem to identify for which $D$-brane configuration $D_i\cup D_f$ and which choice of the two-forms $\alpha_i$ and $\alpha_f$ the cocycle $(\Omega,\alpha_i,\alpha_f)$ is integer-valued. Fortunately enough, if the maximal compact subgroup of $D$ is simple and simply connected, we have the key for solving our problem: we draw the interacting $D$-brane diagrams directly in the double and repeat the discussion of the section 2, using the duality invariant first-order action (25). The action (25) also contains the WZNW term but now the forms $\alpha$ vanish. This means that the pairing of the cocycle $(c,\alpha_F=0,\alpha_{Fd}=0)$ with any relative cycle $\gamma$ is just \begin{equation} \langle (c,\alpha_F,\alpha_{Fd}),\gamma\rangle=\int_{\gamma}c .\end{equation} Recall that we assumed $\pi_1(F)=0$. By the Hurewicz isomorphism, we obtain $H^2(F)=0$ , hence every cycle in the relative singular homology $H_3(R,D_i\cup D_f)$ can be represented by a cycle in $H_3(R)$. This means that what matters is only whether $c$ is the standard integer-valued three-cocycle in the third de Rham cohomology $H^3(D)$ of the Drinfeld double. But it is, because $H^3(D)=H^3(K)$, where $K$ is the simple simply connected maximal compact subgroup of $D$, and it is known that the WZNW three-form restricted to $K$ is the integer-valued cocycle. We find quite appealing that the path integral for the $D$-branes configurations seems to be topologically more easily tractable by using the duality invariant formalism on the Drinfeld double. On the other hand, an account of the local world-sheet phenomena, like a short-distance behaviour, seems to be more difficult when working with the non-manifestly Lorentzian first order Hamiltonian action (25). We plan to study this issue in detail in a near future. \section{Example: $SU(N)$ WZNW model} Now we shall study examples of this general construction of the self-dual WZNW models. Consider the group $SL(N,{\bf C})$ viewed as the real group and the following invariant non-degenerate bilinear form on its algebra\footnote{The normalization of the bilinear form is always such that the resulting action of the $SU(N)$ WZNW model will be properly normalized in order to meet the requirement that the WZNW three-form is the integer-valued cocycle.} \begin{equation} \langle X,Y\rangle = {\rm Im}[(a^*)^2{\rm Tr} XY], \qquad {\rm Im} a^2=4.\end{equation} The group $SL(N,{\bf C})$ equipped with the invariant bilinear form is the Drinfeld double for every choice of the complex parameter $a$ satisfying the normalization constraint in (62). Two isotropic subalgebras ${\cal G}$ and $\ti{\cal G}$ of ${\cal D}$ are all upper and lower triangular matrices respectively with diagonal elements being $\lambda_k a, ~\lambda_k\in {\bf R}$ for ${\cal G}$ and $\tilde\lambda_k ia, ~\tilde\lambda_k\in {\bf R}$ for $\ti{\cal G}$. Obviously, the index $k$ denotes the position on the diagonal and lambdas are constrained by the tracelessness condition. An example of $SL(2,{\bf C})$: \begin{equation} {\cal G}=\left(\matrix{\lambda a&z\cr 0& -\lambda a}\right), \quad \ti{\cal G}=\left(\matrix{\tilde\lambda ia&0\cr \tilde z&-\tilde\lambda ia}\right),\end{equation} where $z,\tilde z$ are arbitrary complex numbers. The dual pair of the $\sigma$-models is encoded in the choice of the half-dimensional subspace ${\cal R}$ of the Lie algebra ${\cal D}$ of the double. We choose ${\cal R}$ to be the $su(N)$ subalgebra of the algebra $sl(N,{\bf C})$. Following our discussion above, it is easy to find the principal fibrations of $SL(N,{\bf C})\equiv D$, corresponding to the algebras ${\cal R}$, ${\cal G}$ and $\ti{\cal G}$. The total space of the bundles is always the double $D$, the fibres are $SU(N)$, $\exp{{\cal G}}\equiv G$ and $\exp{\ti{\cal G}}\equiv \tilde G$ and the bases are $SU(N)\backslash D$, $D/G$ and $D/\tilde G$ respectively. Note that every fiber of all three fibrations can be obtained from the fiber crossing the unit element $e$ of $D$ by either the right (for $SU(N)$) or the left (for $G$ and $\tilde G$) action of some element of the double $D$. In the particular example of the double $SL(N,{\bf C})$, the intersection of a fibre $SU(N)$ with fibres $G$ or $\tilde G$ occurs always precisely at one point. It is not difficult to prove this fact. We already know that the intersection always exists because $SU(N)$ is compact and $SL(N,{\bf C})$ is connected (cf. sec 3). If both fibers $SU(N)$ and $G$ (or $\tilde G$) cross the unit element of the double (which is the intersection point), it is obvious that a non-unit element of $G$ (or $\tilde G$) cannot be a unitary matrix. Thus the intersection is unique in this case. Also an intersection $r$ of the $SU(N)$ fiber crossing the unit element of $D$ with some fiber $G$ must be unique. Indeed, the $G$ fiber can be then written as $rG$ where $r\in SU(N)$. By the left action of $r^{-1}$ the $G$ fiber can be transported to the origin of $D$ where there is just one intersection. Hence we conclude: for our data $D=SL(N,{\bf C}),G,\tilde G$ and ${\cal R}=su(N)$, the both models of the dual pair (16) are the standard $SU(N)$ WZNW models, because the restriction of the bilinear form (62) to ${\cal R}$ is nothing but the standard Killing-Cartan form on $su(N)$. Now we may choose the subgroup $F$ of $D$, which defines the $D$-branes in the double, to be equal to $G$. Thus we have given a concrete meaning to our so far abstract construction. It may be of some interest to provide few explicit formulas for the $SL(2,{\bf C})$ Drinfeld double. The both cosets $D/G$ and $D/\tilde G$ can be identified with the group $SU(2)$. Recall that the space of all $D$-branes corresponding to the choice $F$ is parametrized by the {\it left} coset $F\backslash D$. In our case $F\backslash D$ can also be identified with $SU(2)$, hence a generic $D$-brane ($Fd$) in the double is a set of $SL(2,{\bf C})$ matrices of the form \begin{equation} Fd \equiv \left(\matrix{e^{\lambda a}&z\cr 0&e^{-\lambda a}}\right) \left(\matrix{C&-E^*\cr E&C^*}\right),\end{equation} where $C,E$ are fixed complex numbers satisfying $CC^* +EE^*=1$, $\lambda$ is a real and $z$ a complex number. In order to get the $D$-branes in the cosets $D/G=D/F$ and $D/\tilde G$, we have to project $Fd$ on $SU(2)$ along $G$ and $\tilde G$, respectively: \begin{equation} Fd= \left(\matrix{A&-B^*\cr B&A^*}\right) \left(\matrix{e^{\eta a}&w\cr 0&e^ {-\eta a}}\right), \quad \eta\in {\bf R}, \quad w\in {\bf C} ,\end{equation} \begin{equation} Fd= \left(\matrix{\tilde A&-\tilde B^*\cr \tilde B&\tilde A^*}\right) \left(\matrix{e^{\tilde\eta ia} &0\cr \tilde w&e^{-\tilde\eta ia}}\right), \quad \tilde\eta\in {\bf R}, \quad \tilde w\in {\bf C} .\end{equation} Here again $AA^*+BB^*=1$ and the same constraint is of course true also for $\tilde A$ and $\tilde B$. If $\lambda$ and $z$ vary then $A$ and $B$ sweep a submanifold of $SU(2)$, which is just the $D$-brane in $R$ and $\tilde A$ and $\tilde B$ sweep the dual $D$-brane in $R$. There may occur three qualitatively different possibilities: \noindent 1. Both $C$ and $E$ do not vanish (a generic case). \noindent Then $B\neq 0$ and it is convenient to parametrize $D$ and $B$ as \begin{equation} E=e^{E_1 a} e^{E_2 ia},\qquad B=e^{B_1 a} e^{B_2 ia}, \quad B_i,E_i\in{\bf R} .\end{equation} The original $D$-brane is then a two-dimensional submanifold of $SU(2)$ characterized by the condition \begin{equation} B_2=E_2.\end{equation} The dual $D$-brane is a three dimensional submanifold of $SU(2)$ which is complement of the circle $A=0$. \noindent 2. $C=0$. \noindent The original $D$-brane is the same as in 1. but the dual $D$-brane is just the one-dimensional circle $A=0$. \noindent 3. $E=0$. \noindent The original $D$-brane is a point $A=C$ and the dual $D$-brane is the same as in 1. It is not difficult to compute also the two-form $\alpha$ on the $D$-brane (cf. (38)). For doing this, we have just to know the mapping $k(r)$ from the original $D$-brane in $R$ to $G$ and the dual mapping $\tilde k(r)$ from the dual $D$-brane in $R$ to $\tilde G$ (cf. (35)). We do that for the case 3 choice $C=1$ and $E=0$. The original map $k$ is trivial, since the $D$ brane is just the point $A=1$ but the dual map $\tilde k$ is nontrivial and it reads \begin{equation} \tilde k(A,B)=\left(\matrix{e^{iA_2 a}&0\cr -e^{-A_1 a}B&e^{-iA_2 a}}\right). \end{equation} Here \begin{equation} 0\neq A^*\equiv e^{A_1 a}e^{iA_2 a}, \quad A_1,A_2\in{\bf R}.\end{equation} Now insert (69) in (38) and find \begin{equation} 8\pi\tilde\alpha_f= {\rm Im}[{a^*}^2\{(-BdB^*+B^*dB)a\wedge(A_1+idA_2)+ dB^*\wedge dB -2ia^2 dA_1\wedge dA_2\}].\end{equation} It is easy to compute the exterior derivative of $\tilde\alpha_f$: \begin{equation} 8\pi d\tilde\alpha_f=2{\rm Im}[{a^*}^2 a~ dB^*\wedge dB \wedge d(A_1+iA_2)]= c_R\vert_{D_f}.\end{equation} In words: the exterior derivative of $\alpha_f$ is equal to the restriction of the WZNW three-form $c_R$ on the $D$-brane $D_f$. It may be interesting to remark that in the case of the $SU(N)$ WZNW models there are no topological obstructions in quantizing the model on the topologically trivial open strip world-sheet. Thus, we do not have to lift the $D$-brane configuration to the double in order to make the argument but we can directly proceed at the level of the $D/G\equiv SU(N)$ target. Indeed, choose two different surfaces lying in the same $D$-brane and interpolating between the edges of the `diadem'. Their oriented sum does not have a boundary and it is topologically the two-sphere. If we happen to show that the second homotopy group $\pi_2(D_{i(f)})$ of the $D$-brane $D_{i(f)}$ vanishes then the two interpolating surfaces are homotopically equivalent and there is no ambiguity coming from the WZNW term (cf. sec 2). It is easy to prove that $\pi_2(D_f)$ vanishes if the $D$-brane $D_f$ was obtained by our method of projecting the isotropic surface $Fd$ from the double. Our basic tool is the long exact homotopy sequence \cite{Schw}: \begin{equation} \pi_2(F)=0\to \pi_2(F/H)\to\pi_1(H)\to \pi_1(F)\to \pi_1(F/H)\to \pi_0(H)\to 0=\pi_0(F),\end{equation} which holds for a connected group $F$ and its arbitrary subgroup $H$; note that $\pi_2$ of any Lie group vanishes. Now the $D$-brane on $SU(N)$ is gotten by projection of the surface $Fd$ from the double to $D/G$. This means that topologically it can be identified with the coset $F/dGd^{-1}\cap F\equiv F/H$. We observe that in our $SL(N,{\bf C})$ context the group $F$ can be topologically identified with its algebra ${\cal F}$ because the usual exponential mapping $\exp{{\cal F}}=F$ is one-to-one. So it is one-to-one for any its connected subgroup including the unity component of $H$. Hence $\pi_1(H)=0$ and since $\pi_1(F)=0$, from the sequence (73) we conclude that $\pi_2$ of the $D$-brane in $SU(N)$ vanishes. We should mention that from the exact sequence (73) it also follows that $\pi_1(D_f)=\pi_0(H)$. In general, for our $SU(N)$ case the group $H$ is not connected. This means that, strictly speaking, the diadem in our argument must be equivalent to the zero element of $H_2(R,D_f)$, or, in other, words it must be a relative two-boundary. If the diadem is a non-trivial relative two-cycle we use the results of section 2 and evaluate its contribution by choosing the extension $\tilde f:Z\to U$ of the homomorphism $f:B\to U$.
proofpile-arXiv_065-462
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} \begin{figure} \vspace*{13pt} \vspace*{6.7truein} \special{psfile=figa.ps voffset= 240 hoffset= -40 hscale=50 vscale=50 angle = 0} \special{psfile=figb.ps voffset= 240 hoffset= 220 hscale=50 vscale=50 angle = 0} \special{psfile=figc.ps voffset= -40 hoffset= 90 hscale=50 vscale=50 angle = 0} \caption{Histogram of the number of models that yield a particular prediction for $m_{\nu_{\mu}}^2- m_{\nu_{e}}^2$ assuming (a) small angle and (b) large angle solution to solar neutrino problem. In (c) we solve the solar neutrino problem via small angle $e$--$\tau$ oscillations and check whether this is compatible with the LSND result. } \label{fig} \end{figure} In the Standard Model of elementary particles (SM) both lepton number ($L$) and baryon number ($B$) are conserved due to an accidental symmetry, {\sl i.e.} there is no renormalizable, gauge-invariant term that would break the symmetry. In the minimal supersymmetric extension of the SM (MSSM) the situation is different. Due to a the variety of scalar partners the MSSM allows for a host of new interactions many of which violate $B$ or $L$. Since neither $B$ nor $L$ violation has been observed in present collider experiments these couplings are constrained from above. More constraints arise from neutrino physics or cosmology. Thus, all lepton and baryon number violating interaction are often eliminated by imposing a discrete, multiplicative symmetry called $R$-parity,\cite{r-parity} $R_p \equiv (-1)^{2S+3B+L}$, where $S$ is the spin. One very attractive feature of $R_p$ conserving models is that the lightest supersymmetric particle (LSP) is stable and a good cold dark matter candidate.\cite{cdm} However, while the existence of a dark matter candidate is a very desirable prediction, it does not prove $R_p$ conservation and one should consider more general models. Here, we will investigate the scenario where $R_p$ is broken explicitly via the terms\cite{suzuki} $W = \mu_i L_i H$, where $H$ is the Higgs coupling to up-type fermions and $L_i$ ($i = 1,2,3$) are the left-handed lepton doublets. Clearly, these Higgs-lepton mixing terms violate lepton-number. As a result, majorana masses will be generated for one neutrino at tree-level and for the remaining two neutrinos at the one-loop level. These masses were calculated in the frame-work of minimal supergravity in ref.~\citenum{npb} and the numerical results will be briefly summarized here. There are three $R_P$ violating parameters which can be used to fix 1) the tree-level neutrino mass, 2) the $\mu$--$\tau$ mixing angle and 3) the $e$--$\mu$ mixing angle. The question of whether e.g. the solar\cite{solarn} and the atmospheric\cite{atmosphericn} neutrino puzzle can be solved simultaneously depends on the prediction of $m_{\nu_\mu}^2-m_{\nu_e}^2$. In fig.~1 we have scanned the entire SUSY parameter space consisting of the Higgsino (gaugino) mass parameter, $\mu$ ($m_{1/2}$), the trilinear scalar interaction parameter $A_0$, and the ratio of Higgs VEVs, $\tan\beta$. The universal scalar mass parameter $m_0$ is fixed by minimizing the potential. Plotted is the number of models yielding a particular prediction for $m_{\nu_\mu}^2-m_{\nu_e}^2$ for (a) sin$^2 2 \theta_{e \nu_\mu} = 0.008$ and (b) sin$^2 2 \theta_{e \nu_\mu} = 1$. We fix $m_{\nu_\tau}=0.1$~eV and sin$^2 2 \theta_{\mu \nu_\tau} = 1$ in order to solve the atmospheric neutrino problem. We see that both long wave-length oscillation (LWO)\cite{lwo} ($m_{\nu_\mu}^2-m_{\nu_e}^2=10^{-10}$~eV$^2$) and MSW effect\cite{msw-effect} ($m_{\nu_\mu}^2-m_{\nu_e}^2=10^{-5}$~eV$^2$) can be accommodated. In fig.~1(c) we solve the solar neutrino problem via $e$--$\tau$ oscillations and we fix sin$^2 2 \theta_{e \nu_\mu} = 0.004$ in order to accommodate the LSND result.\cite{lsnd} We see that most models are already ruled out by collider constraints and even more by dark matter (DM) constraints. However, a very small (but non-zero) number of models yields a prediction compatible with the LSND result (the dotted line is lower limit of LSND). \noindent{\bf Acknowledgements} This work was supported in parts by the DOE under Grants No. DE-FG03-91-ER40674 and by the Davis Institute for High Energy Physics. \vskip0.3cm \noindent{\bf References}
proofpile-arXiv_065-463
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section*{Introduction} To define quasilocal energy in general relativity, one can begin with a suitable action functional for the time history ${\cal M}$ of a spatially bounded system $\Sigma$. Here ``suitable'' means that in the associated variational principle the induced metric on the time history ${\cal T}$ of the system boundary $B = \partial\Sigma$ is fixed. In particular, this means that the lapse of proper time between the boundaries of the initial and final states of the system $\Sigma$ must be fixed as boundary data. The quasilocal energy (QLE) is then defined as minus the rate of change of the classical action (or Hamilton-Jacobi principal function) corresponding to a unit increase in proper time.\cite{BY,Lau} So defined, the QLE is a functional on the gravitational phase space of $\Sigma$, and is the value of the gravitational Hamiltonian corresponding to unit lapse function and zero shift vector on the system boundary $B$. Although other definitions of quasilocal energy have been proposed (see, for example, the references listed in \cite{BY}), the QLE considered here has the key property, which we consider crucial, that it plays the role of internal energy in the thermodynamical description of coupled gravitational and matter fields.\cite{thermo} In this paper we define the energy of a perfectly isolated system at a given retarded time as the suitable limit of the quasilocal energy $E$ for the partial system enclosed within a finite topologically spherical boundary.\footnote{Hecht and Nester have also considered energy-momentum (and ``spin'') at null infinity (for a class of generally covariant theories including general relativity) {\em via} limits of quasilocal Hamiltonian values.\cite{HechtNester} Their treatment of energy-momentum is based on a differential-forms version of canonical gravity, often referred to as the ``covariant canonical formalism." For pure {\sc bms} translations our results are in accord with those found by Hecht and Nester, although at the level of {\em general} supertranslations they differ. We provide a careful analysis of the zero-energy reference term (necessary for the QLE to have a finite limit at null infinity), and this analysis is intimately connected with our results concerning general supertranslations.} For our choice of asymptotic reference frame the energy that we compute equals what is usually called the Bondi-Sachs mass.\cite{Goldberg,review} As we shall see, our asymptotic reference frame defines precisely that infinitesimal generator of the Bondi-Metzner-Sachs ({\sc bms}) group corresponding to a pure time translation.\cite{Sachs,review,PenroseRindler} We also show that in the same null limit the lapse-arbitrary, shift-zero Hamiltonian boundary value defines a physically meaningful element in the space dual to supertranslations. This dual space element, it turns out, coincides with the ``supermomentum" discussed by Geroch.\cite{Geroch} Our results are then specialized to an expression for the full Bondi-Sachs four-momentum in terms of Hamiltonian values. It is already known that when $B$ is the two-sphere at spacelike infinity, the quasilocal and Arnowitt-Deser-Misner\cite{ADM} notions of energy-momentum agree.\cite{BY,thermo} Our results therefore indicate that the quasilocal formalism provides a unified Hamiltonian framework for describing the standard notions of gravitational energy-momentum in asymptopia. Before turning to the technical details, let us first present a short overview of our approach. Consider a spacetime ${\cal M}$ which is asymptotically flat at future null infinity ${\cal I}^{+}$ and a system $(w,R,\theta,\phi)$ of Bondi coordinates thereon.\cite{review} The retarded time $w$ labels a one-parameter family of outgoing null hypersurfaces ${\cal N}(w)$. The coordinate $R$ is a luminosity parameter (areal radius) along the outgoing null-geodesic generators of the hypersurfaces ${\cal N}(w)$. The Bondi coordinate system also defines a two-parameter family of topologically spherical two-surfaces $B(w,R)$. It suits our purposes to consider only a single null hypersurface of the family ${\cal N}(w)$, say ${\cal N}(w_{*})$, the one determined by setting $w$ equal to an {\em arbitrary} constant $w_{*}$. The collection $B(w_{*},R)$ of two-surfaces foliates ${\cal N}(w_{*})$, and in the $R\to\infty$ limit these two-surfaces converge on an infinite-radius {\em round} sphere $B(w_{*},\infty)$. To streamline the presentation, we refer to our generic null hypersurface simply as $\cal N$; and we use the plain letter $B$ to denote both the $\cal N$-foliating collection $B(w_{*},R)$ and a single generic two-surface of this collection. Now, should we desire a more general ${\cal N}$-foliating collection of two-surfaces, we could, of course, introduce a new radial coordinate $\bar{R}$. For a fixed retarded time $w = w_{*}$ the new two-surfaces would then arise as level surfaces of constant $\bar{R}$. However, we shall not consider such a new radial coordinate, because the new two-surfaces would not necessarily converge towards a round sphere in the asymptotic limit. At any rate, we could handle such an additional kinematical freedom, were it present, by assuming that along each outgoing null ray $\bar{R}$ approached $R$ at a sufficiently fast rate in the asymptotic limit. Our first goal is to compute the QLE within a two-surface $B$ in the limit as $B$ approaches a spherical cut of ${\cal I}^{+}$ along the null surface $\cal N$, and to show that this result coincides with the Bondi-Sachs mass: \begin{equation} M_{_{BS}}(w_{*}) = \lim_{R \rightarrow \infty} \int_{B(w_{*},R)} {\rm d}^{2}x \sqrt{\sigma} \varepsilon\, . \label{limit} \end{equation} Here $\varepsilon = (k - k |^{\scriptscriptstyle {\rm ref}})/\kappa$ is the quasilocal energy surface density with $\kappa = 8\pi$ (in geometrical units) and $\sigma$ is the determinant of the induced metric on $B$. Recall that $k$ denotes the mean curvature of $B$ as embedded in some {\em spacelike} spanning three-surface $\Sigma$. Since both $B$ and $\Sigma$ are embedded in the {\em physical} spacetime ${\cal M}$, we sometimes use the notation $\varepsilon |^{\scriptscriptstyle {\rm phy}} = k/\kappa$. Also recall that $k |^{\scriptscriptstyle {\rm ref}}$ denotes the mean curvature of a surface which is isometric to $B$ but which is embedded in a three-dimensional {\em reference} space different than $\Sigma$. Here we choose the reference space to be flat Euclidean space $E^{3}$, i.\ e.\ we assign a flat three-slice of Minkowski spacetime the zero value of energy.\cite{BY} Although a definition of the zero-energy reference in terms of flat space is neither always essential nor appropriate\cite{Brown}, it is the correct choice for the analysis of this paper. In order to define $k$, we must select such a three-surface spanning $B$ for each $R$ value. (For a single $B$ many different spanning three-surfaces will determine the same $k$. In fact, $k$ is determined solely by $B$ and a timelike unit vector field $u^\mu$ on $B$, which can be considered as the unit normal of a slice $\Sigma$. Thus, the continuation of $\Sigma$ away from $B$ is not needed; moreover, such a continuation of $\Sigma$ might not be defined throughout the interior of ${\cal M}$. Therefore, though we speak of choosing a $\Sigma$ three-surface to span $B$ for each $R$ value, we are really fixing only a timelike unit normal vector field at $B$.) For generality, we leave the choice of spanning three-surface $\Sigma$ essentially arbitrary at each $R$ value, but we do enforce a definite choice asymptotically. Heuristically, as $R \rightarrow \infty$ the $\Sigma$ three-surface spanning $B$ approaches an asymptotic three-surface $\Sigma_{\infty}$ which spans a round infinite-radius spherical cut of ${\cal I}^{+}$ (see the figure). Our construction is, as expected, sensitive to the choice of asymptotic three-surface $\Sigma_{\infty}$. Said another way, the QLE depends on the fleet of observers at $B$ whose four-velocities are orthogonal to the spanning three-surface at $B$. Therefore, one expects {\em a priori} the expression on the right-hand side of (\ref{limit}) to depend on the choice of asymptotic fleet associated with the two-sphere at ${\cal I}^{+}$. The asymptotic fleet we choose corresponds to a pure {\sc bms} time translation: each member of the asymptotic fleet rides along $\partial/\partial w$. Note that, although $\partial/\partial w$ is everywhere timelike in ${\cal M}$ (at least in the relevant exterior regions), the {\em extension} of $\partial/\partial w$ to ${\cal I}^{+}$ in a conformal completion $\hat{\cal M}$ of the physical spacetime ${\cal M}$ is in fact a null vector which lies in ${\cal I}^{+}$. (While we occasionally find it clarifying to make reference to the concept of a conformal completion, we do not explicitly use conformal completions in this paper.) Therefore, heuristically, one should envision $\Sigma_{\infty}$ as a spacelike slice which becomes null asymptotically (see the figure). This paper is organized as follows. In a preliminary section we write down the familiar Bondi-Sachs form\cite{Sachs,Chrusciel_et_al} of the spacetime metric as well as asymptotic expansions for the associated metric coefficients. We also introduce on $\cal M$ two future-pointing null vector fields $k^{\mu}$ and $l^{\mu}$ (do not confuse $k^{\mu}$ with the mean curvature $k$). Both vector fields point everywhere normal to our collection $B(w,R)$ of two-surfaces, and $k_{\mu} l^{\mu} = - 1$. Next, we construct on ${\cal N}$ a timelike vector field $u^{\mu} := \frac{1}{2} k^{\mu} + l^{\mu}$ (equality restricted to ${\cal N}$), which in our analysis will {\em define} for each $B$ along $\cal N$ a spacelike spanning three-surface $\Sigma$. In $\S$II we use the $\Sigma$ three-surfaces determined by $u^{\mu}$ to define an unreferenced energy surface density $\varepsilon |^{\scriptscriptstyle {\rm phy}} = k/\kappa$ for each $B$ slice of $\cal N$ and then examine the asymptotic limit of $k/\kappa$. In $\S$III we consider the asymptotic expression for the flat-space reference density $\varepsilon |^{\scriptscriptstyle {\rm ref}} = k |^{\scriptscriptstyle {\rm ref}}/\kappa$, but give the derivation of this expression in the Appendix. Next, we assemble the results of the previous two sections and prove the main claim (\ref{limit}). In $\S$IV we examine the ``smeared energy surface density,'' which is the Hamiltonian value corresponding to an arbitrary supertranslation. We then specialize our result for the smeared energy surface density to express the full Bondi-Sachs four-momentum in terms of Hamiltonian values. In $\S$V we examine the smeared energy surface density {\em via} the spin-coefficient formalism, and show that it equals the ``supermomentum'' of Geroch\cite{Geroch} as written by Dray and Streubel.\cite{DrayStreubel} The Appendix is devoted to a detailed analysis of the reference term. \section{Preliminaries} In terms of a Bondi coordinate system the metric of our asymptotically flat spacetime ${\cal M}$ takes the standard form\cite{Sachs,Chrusciel_et_al} \begin{equation} g_{\mu\nu} {\rm d}x^{\mu} {\rm d}x^{\nu} = - UV {\rm d}w^{2} - 2 U {\rm d}w {\rm d}R + \sigma_{ab} ({\rm d}x^{a} + W^{a}{\rm d}w) ({\rm d}x^{b} + W^{b}{\rm d}w){\,} , \label{spacetimemetric} \end{equation} where $a,b$ are $B$ indices running over $\theta,\phi$. We assume the following expansions for the various metric coefficients above:\footnote{Up to the $\Delta$ remainder terms, our expansions in the radial coordinate $R$ coincide with those given by Sachs; however, we do not assume that the $\Delta$ remainder terms are necessarily expandable in powers of inverse $R$, an assumption which would be tantamount to what Sachs calls the ``outgoing radiation condition.''\cite{Sachs} Recently, Chru\'{s}ciel {\em et al.}~have shown that ``polyhomogeneous'' expansions in terms of $R^{-i}\log^{j}\!R$ also provide a consistent framework for solving the characteristic initial value problem of the Bondi-Sachs type.\cite{Chrusciel_et_al} They argue that the so-called outgoing radiation condition is overly restrictive.} \begin{eqnarray} V & = & 1 - 2{\rm m} R^{-1} + \Delta_{V} \label{Sachs} \eqnum{\ref{Sachs}a} \\ U & = & 1 - {\textstyle \frac{1}{2}}(X^{2} + Y^{2}) R^{-2} + \Delta_{U} \eqnum{\ref{Sachs}b}\\ W^{\theta} & = & (2 X\cot \theta + X_{,\theta} + Y_{,\phi} \csc\theta)R^{-2} + \Delta_{W^{\theta}} \nonumber \\ W^{\phi} & = & \csc\theta\left(2 Y \cot\theta + Y_{,\theta} - X_{,\phi}\csc\theta\right)R^{-2} + \Delta_{W^{\phi}} \eqnum{\ref{Sachs}c} \\ \sigma_{ab} & = & R^{2}\delta_{ab} + \left[(2 X)\theta_{,a}\theta_{,b} + (4 Y \sin\theta)\theta_{,(a} \phi_{,b)} - (2 X \sin^{2}\theta) \phi_{,a} \phi_{,b}\right]R + \Delta_{\sigma_{ab}}{\,} . \eqnum{\ref{Sachs}d} \addtocounter{equation}{1} \end{eqnarray} Here $X(w,\theta,\phi)$ and $Y(w,\theta,\phi)$ are respectively the real and imaginary parts of the asymptotic shear $c = X + {\rm i} Y$, ${\rm m}(w,\theta,\phi)$ is the all-important {\em mass aspect}, $\delta_{ab}$ is the metric of a unit-radius round sphere, and commas denote partial differentiation. In the Appendix we examine the form of the two metric $\sigma_{ab}$ in more detail. Remainder terms, denoted by the $\Delta$ symbol, always fall off faster (or have slower growth, as the case may be) than the terms which precede them. For instance, $\Delta_{V}$ denotes a term which falls off {\em faster} than $O(R^{-1})$. Introduce the future-directed null covector field $k_{\mu} = - e^{\eta} \nabla_{\mu} w$, where the scalar function $\eta = \eta(w, R , \theta , \phi)$ is a point-dependent boost parameter. The null covector $k_{\mu}$ is orthogonal to the spheres $B(w,R)$, and the function $\eta$ gives us complete freedom in choosing the extent of $k_{\mu}$ at each point of any $B$ two-surface. We shall find it necessary later to assume that $\eta$ falls off {\em faster} than $1/\sqrt{R}$ on every outgoing ray. Also define another future-directed null vector field $l^{\mu}$ which is orthogonal to the $B(w,R)$ and normalized so that $k_{\mu} l^{\mu} = - 1$. As one-forms these null normals are \begin{eqnarray} k_{\mu} {\rm d}x^{\mu} & = & - e^{\eta} {\rm d}w \label{bassae} \eqnum{\ref{bassae}a} \\ l_{\mu} {\rm d}x^{\mu} & = & - e^{-\eta} U {\rm d}R - {\textstyle \frac{1}{2}} e^{-\eta} UV {\rm d}w\, , \eqnum{\ref{bassae}b} \addtocounter{equation}{1} \end{eqnarray} while as vector fields they are \begin{eqnarray} k^{\mu} \partial/\partial x^{\mu} & = & e^{\eta} U^{-1} \partial/\partial R \label{vectors} \eqnum{\ref{vectors}a}\\ l^{\mu} \partial/\partial x^{\mu} & = & e^{-\eta} \partial/\partial w - {\textstyle \frac{1}{2}} e^{-\eta} V \partial/\partial R - e^{-\eta} W^{a} \partial/\partial x^{a} \eqnum{\ref{vectors}b} \, . \addtocounter{equation}{1} \end{eqnarray} Now define $u^\mu := \frac{1}{2} k^\mu + l^\mu$ and $n^\mu := \frac{1}{2} k^\mu - l^\mu$ along ${\cal N}$ as the timelike and spacelike unit normals of the $B$ two-surfaces. For each slice $B$ of the null hypersurface ${\cal N}$, the normals $u^\mu$ and $n^\mu$ determine a spanning spacelike three-surface $\Sigma$. As mentioned previously, the three-surface $\Sigma$ is not unique and, moreover, need not be defined throughout ${\cal M}$. Indeed, there is no guarantee that $u^\mu$ as defined is even surface-forming. (That is, in general $u_\mu$ does not satisfy the Fr\"obinius condition $u_{[\alpha}\nabla_{\mu} u_{\nu]}=0$.) Nevertheless, our construction provides us with what we need: a unit timelike vector $u^{\mu}$ orthogonal to $B$. We can therefore obtain an unreferenced energy surface density $k/\kappa$ which is the same for any slice or partial slice $\Sigma$ that contains $B$ and has timelike unit normal which agrees with $u^\mu$ at $B$. Our construction implies \begin{equation} u^{\mu}\partial/\partial x^{\mu} \rightarrow \partial/\partial w \end{equation} on each ray as $R \rightarrow \infty$. Now, the standard realization of the {\sc bms}-group Lie algebra (as vector fields on future null infinity) identifies the {\em extension} of $\partial/\partial w$ to ${\cal I}^{+}$ (in a conformal completion $\hat{\cal M}$ of ${\cal M}$) with a pure time translation.\cite{Sachs,review} Therefore, asymptotically, our fiducial surface $\Sigma_{\infty}$ determines precisely the pure time-translation generator of the {\sc bms} group. We do {\em not} claim that $u^{\mu}$ generates an ``infinitesimal asymptotic symmetry transformation'' in the sense of Sachs\cite{Sachs}, {\it i.~e.~}that the various coefficients associated with the transformed metric $g_{\mu\nu} + 2\nabla_{(\mu} u_{\nu)}$ satisfy the fall-off conditions (\ref{Sachs}); however, this is unimportant for our construction. \section{Computation of the quasilocal energy surface density} We now turn to the task of calculating an expression for the unreferenced quasilocal energy surface density $\varepsilon |^{\scriptscriptstyle {\rm phy}} = k/\kappa$. Our starting point is the definition $k := - \sigma^{\mu\nu} \nabla_{\mu} n_{\nu}$, where the two-metric $\sigma^{\mu\nu} = g^{\mu\nu} + 2k^{(\mu} l^{\nu)}$ serves as the projection operator into $B$. We find it convenient to write\footnote{Note that the definition of $k$ does not depend on how $n^{\mu}$ is extended off $B$.} $k = 2\mu + \rho$, where in the standard notation of the spin-coefficient formalism\cite{PenroseRindler} $- \mu$ and $\rho$ are, respectively, the expansions associated with the inward null normal and outward null normal to $B$. These are given by the formulae \begin{eqnarray} \mu & = & {\textstyle \frac{1}{2}} \sigma^{\mu\lambda} \nabla_{\mu} l_{\lambda} = {\textstyle \frac{1}{2}}(\nabla_{\mu} l^{\mu} + k^{\nu} l^{\mu} \nabla_{\mu} l_{\nu})\label{spin} \eqnum{\ref{spin}a} \\ \rho & = & - {\textstyle \frac{1}{2}} \sigma^{\mu\lambda}\nabla_{\mu} k_{\lambda} = - {\textstyle \frac{1}{2}}(\nabla_{\mu} k^{\mu} + l^{\nu} k^{\mu} \nabla_{\mu} k_{\nu})\, . \eqnum{\ref{spin}b} \addtocounter{equation}{1} \end{eqnarray} As a technical tool, it proves convenient to introduce fiducial vector fields $\hat{k}^{\mu}$ and $\hat{l}^{\mu}$ determined from (\ref{vectors}) by setting $\eta = 0$ on $\cal N$. {}From the middle expressions above, it is obvious that $\mu = e^{-\eta} \hat{\mu}$ and $\rho = e^{\eta} \hat{\rho}$, where the easier-to-calculate expressions $\hat{\mu}$ and $\hat{\rho}$ are built exactly as in (\ref{spin}) but with the fiducial null normals $\hat{k}^{\mu}$ and $\hat{l}^{\mu}$. Therefore, we may assume that $\eta = 0$ while calculating the spin coefficients in (\ref{spin}) and then simply multiply the $\eta = 0$ results by the appropriate factor to get the correct general expressions. Let us sketch the calculation. First, from (\ref{bassae}a) with $\eta = 0$ note that $\hat{k}^{\mu} \nabla_{\mu} \hat{k}_{\nu} = 0$, because $\hat{k}_{\nu}$ is a gradient. Next, using both expressions (\ref{bassae}) with $\eta = 0$, one can work the second term inside the parenthesis of (\ref{spin}a) into the form $\hat{k}^{\nu} \hat{l}^{\mu} \nabla_{\mu} \hat{l}_{\nu} = - U^{-1} \hat{l}^{\mu} \nabla_{\mu} U + \frac{1}{2} U \hat{k}^{\mu} \nabla_{\mu} V$. Finally, one writes the covariant-divergence terms as ordinary divergences; for example, $\nabla_{\mu} \hat{k}^{\mu} = (- g)^{-1/2} \partial_{\mu}(\sqrt{-g} \hat{k}^{\mu})$, where the square root of (minus) the determinant of the spacetime metric is $\sqrt{- g} = U\sqrt{\sigma}$. Following these steps and multiplying by the appropriate boost factors at the end of the calculation, one finds \begin{eqnarray} \mu & = & {\textstyle \frac{1}{4}} e^{-\eta} \sigma^{-1} \dot{\sigma} - {\textstyle \frac{1}{8}} e^{-\eta} V \sigma^{-1}\sigma' - {\textstyle \frac{1}{2}} e^{-\eta} \delta_{a} W^{a} \label{mu1} \eqnum{\ref{mu1}a} \\ \rho & = & - {\textstyle \frac{1}{4}} e^{\eta} U^{-1} \sigma^{-1}\sigma'{\,} . \eqnum{\ref{mu1}b} \addtocounter{equation}{1} \end{eqnarray} Here the over-dot denotes partial differentiation by $\partial/\partial w$, the prime denotes partial differentiation by $\partial/\partial R$, and $\delta_{a}$ denotes the $B$ covariant derivative. Since $R$ is an areal radius, we may take $\sigma = R^{4}\sin^{2}\theta$ [see the form of the $B$ metric given in (\ref{Bmetric})]. Therefore, we obtain the compact expressions \begin{eqnarray} \mu & = & - {\textstyle \frac{1}{2}} e^{-\eta} V R^{-1} - {\textstyle \frac{1}{2}} e^{-\eta} \delta_{a} W^{a} \label{mu2} \eqnum{\ref{mu2}a} \\ \rho & = & - e^{\eta} U^{-1} R^{-1}\, . \eqnum{\ref{mu2}b} \addtocounter{equation}{1} \end{eqnarray} Adding twice (\ref{mu2}a) to (\ref{mu2}b), we arrive at our desired expression \begin{equation} k = - ( e^{-\eta} V + e^{\eta} U^{-1}) R^{-1} - e^{-\eta} \delta_{a} W^{a}\, , \end{equation} which has the asymptotic form \begin{equation} k = - 2R^{-1} + 2{\rm m}(w_{*},\theta,\phi) R^{-2} - \delta_{a} W^{a} + \Delta_{k}\, . \label{kexpansion} \end{equation} Note that we have chosen not to expand the $O(R^{-2})$ pure divergence term $-\delta_{a}W^{a}$. Our assumption about the fall-off of $\eta$ ensures that a term $-\eta^{2}/R$ which appears in the asymptotic expression for $k$ can be swept into $\Delta_{k}$. \section{The Bondi-Sachs mass} Write the total quasilocal energy as $E = E |^{\scriptscriptstyle {\rm phy}} - E |^{\scriptscriptstyle {\rm ref}}$, with the total {\em unreferenced} quasilocal energy $E |^{\scriptscriptstyle {\rm phy}}$ taken as \begin{equation} E |^{\scriptscriptstyle {\rm phy}} = \frac{1}{\kappa} \int_{B(w_{*},R)} {\rm d}^{2}x \sqrt{\sigma} k\, . \label{unreferencedE1} \end{equation} Plugging the expansion (\ref{kexpansion}) into the above expression, using $\sqrt{\sigma} = R^{2}\sin\theta$ for our choice of coordinates, and integrating term-by-term, one finds \begin{equation} E |^{\scriptscriptstyle {\rm phy}} = - R + M_{_{BS}}(w_{*}) + \Delta_{E |^{\scriptscriptstyle {\rm phy}}}\, . \label{unreferencedE2} \end{equation} Here the Bondi-Sachs mass associated with the $w = w_{*}$ cut of ${\cal I}^{+}$ is the two-surface average of the mass aspect evaluated at $w = w_{*}$,\cite{Sachs,Goldberg} \begin{equation} M_{_{BS}}(w_{*}) = \frac{2}{\kappa} \int {\rm d}\Omega\, {\rm m}(w_{*},\theta,\phi)\, . \end{equation} We use the notation $\int {\rm d}\Omega := \int^{\pi}_{0} {\rm d}\theta \int^{2\pi}_{0}{\rm d}\phi\sin\theta$ to denote proper integration over the unit sphere (which is identified with a spherical cut of ${\cal I}^{+}$). In passing from (\ref{unreferencedE1}) to (\ref{unreferencedE2}), we have made an appeal to Stokes' theorem to show that the ``dangerous'' $O(R^{0})$ term that arises from proper integration over the pure-divergence term $\delta_{a} W^{a}$ in (\ref{kexpansion}) does indeed vanish. Hence, this term does not contribute to the Bondi-Sachs mass and does not spoil the result (\ref{unreferencedE2}). The reference point contribution to the energy is \begin{equation} - E |^{\scriptscriptstyle {\rm ref}} = - \frac{1}{\kappa} \int_{B(w_{*},R)} {\rm d}^{2}x \sqrt{\sigma} k |^{\scriptscriptstyle {\rm ref}}\, , \label{referencedE1} \end{equation} where the asymptotic expression for $k |^{\scriptscriptstyle {\rm ref}}$ must be determined from the specific asymptotic form (\ref{Bmetric}) of the Sachs two-metric. We present this calculation in the Appendix. The result is \begin{equation} - E |^{\scriptscriptstyle {\rm ref}} = R + 0 \cdot R^{0} + \Delta_{E |^{\scriptscriptstyle {\rm ref}}}\, . \label{referencedE2} \end{equation} Note the absence of an $O(R^{0})$ term in $E |^{\scriptscriptstyle {\rm ref}}$. The result (3.5) has just the right form, in that it removes the part of $E |^{\scriptscriptstyle {\rm phy}}$ which becomes singular as $R \rightarrow \infty$ but does not itself contribute to the mass. Therefore the total quasilocal energy for large $R$ is \begin{equation} E = \int_{B(w_{*},R)} {\rm d}^{2}x \sqrt{\sigma} \varepsilon = M_{_{BS}}(w_{*}) + \Delta_{E}\, . \end{equation} This is the energy of the gravitational and matter fields associated with the spacelike three-surface $\Sigma$ which spans a $B$ slice of $\cal N$ and which tends toward $\Sigma_{\infty}$. Our main claim (\ref{limit}) follows immediately from (3.6). \section{Smeared Energy Surface Density} Consider the expression $H_{B}$ for the on-shell value of the gravitational Hamiltonian appropriate for a spatially bounded three-manifold $\Sigma$, subject to the choice of a vanishing shift vector at the boundary $\partial\Sigma = B$: \begin{equation} H_{B} = \int_{B} {\rm d}^{2}x \sqrt{\sigma} N \varepsilon {\,} . \label{boundaryH} \end{equation} We refer to $H_B$ as the smeared energy surface density. Addition of this boundary term to the smeared Hamiltonian constraint ensures that as a whole the sum is functionally differentiable.\cite{BY} In this section we consider the $R\to\infty$ limit of $H_B$ along the null hypersurface $\cal N$ in exactly the same fashion as we considered the limit (1.1) of the quasilocal energy previously. Before evaluating $\lim_{R \rightarrow \infty} H_{B(w_{*},R)}$, let us discuss its physical significance. Consider a particular spherical cut $B(w_{*},\infty)$ of ${\cal I}^{+}$. A general {\sc bms} supertranslation pushes $B(w_{*},\infty)$ forward in retarded time $w$ in a general angle-dependent fashion. As is well-known, the infinitesimal generator corresponding to such a supertranslation has the form $\left. \alpha{\,} \partial/\partial w\right|_{{\cal I}^{+}}$, where $\alpha(\theta,\phi)$ is any twice differentiable function of the angular coordinates.\cite{Sachs} As we have seen, $\partial/\partial w$ is heuristically the hypersurface normal $u^{\mu}$ at $B(w_{*},\infty)$ of an asymptotic spanning three-surface $\Sigma_{\infty}$. In other words, each member of the fleet of observers at $B(w_{*},\infty)$ rides along $\partial/\partial w$. Therefore, again heuristically, the on-shell value of the Hamiltonian generator of a general {\sc bms} supertranslation is \begin{equation} \int_{B(w_{*},\infty)}{\rm d}^{2}x \sqrt{\sigma} \alpha \varepsilon\, . \end{equation} This symbolic expression coincides with the $R\to\infty$ limit of the smeared energy surface density (4.1), where we set $\alpha(\theta,\phi) := \lim_{R \rightarrow \infty} N(R,\theta,\phi)$ (suitable fall-off behavior for $N$ is assumed). Thus, $\lim_{R \rightarrow \infty} H_{B(w_{*},R)}$ defines a physically meaningful element in the dual space of general supertranslations. In this respect it is like the ``supermomentum" of Geroch.\cite{Geroch} In the next section we show explicitly that, in fact, $\lim_{R \rightarrow \infty} H_{B(w_{*},R)}$ is precisely Geroch's ``supermomentum." Note, however, that it might be better to call such an expression the ``superenergy,'' as it arises entirely from the ``energy sector'' of the Hamiltonian's boundary term (that is, the sector with vanishing shift vector) but also incorporates the ``many-fingered'' nature of time (that is, an arbitrary lapse function). Let us now evaluate the $R\to\infty$ limit of the smeared energy surface density $H_B$. As we have stated, there is no $O(R^{0})$ contribution to $E |^{\scriptscriptstyle {\rm ref}}$. The absence of this contribution stems from the fact that the two-sphere average of the coefficient ${}^{(2)}\!k |^{\scriptscriptstyle {\rm ref}}$ of the $O(R^{-2})$ piece of the reference term $k |^{\scriptscriptstyle {\rm ref}}$ vanishes. As spelled out in the Appendix, this fact follows directly from an equation governing the required isometric embedding of $B$ into Euclidean three-space. Moreover, as seen in Section 2, the coefficient ${}^{(2)}\! k$ of the $O(R^{-2})$ piece of the physical $k$ is not solely twice the mass aspect but also contains a unit-sphere divergence term. Now, in the present case $\varepsilon = (k - k |^{\scriptscriptstyle {\rm ref}})/\kappa$ is smeared against a function $N$, so one might worry that the limit is spoiled in some way by the presence of the smearing function. However, as we now show, {\em for solutions of the field equations}, the {\em unintegrated} expression $4\pi R^{2}\varepsilon$ is precisely the mass aspect of the system in the $R \rightarrow \infty$ limit. This striking result rests on an exact cancellation between ${}^{(2)}\!k |^{\scriptscriptstyle {\rm ref}}$ and the aforementioned unit-sphere divergence part of ${}^{(2)}\!k$. With the machinery set up in the previous sections and the Appendix [see in particular equations (\ref{kexpansion}) and (A2)], we find the following limit: \begin{equation} \lim_{R \rightarrow \infty} {\textstyle \frac{1}{2}}\kappa R^{2} \varepsilon = {\rm m} (w_{*},\theta,\phi) - {\textstyle \frac{1}{2}} \left[\csc\theta\, \partial_{a} (\sin\theta {}^{(2)}\! W^{a}) -{\textstyle \frac{1}{2}} {}^{(3)}\!{\cal R} \right] \, . \label{limk-k0} \end{equation} Here we set $\kappa = 8\pi$ (in geometrical units) and use ${}^{(3)}\!{\cal R}$ to denote the coefficient of the $O(R^{-3})$ piece of the $B$ Ricci scalar. Also, the coefficients ${}^{(2)}\! W^{a}$ of the leading $O(R^{-2})$ pieces of $W^{a}$ are listed in (1.2c). Inspection of (\ref{Sachs}c,d) shows that ${}^{(2)}\!W^{a}$ is expressed in terms of the same functions, $X$ and $Y$, that appear in the $O(R^{-1})$ piece of $\sigma_{ab}/R^{2}$. Furthermore, a short calculation with the $B$ metric shows that for these solutions ${}^{(3)}\! {\cal R}$ may be expressed in terms of ${}^{(2)}\! W^{a}$ as follows: \begin{equation} - {\textstyle \frac{1}{2}} {}^{(3)}\! {\cal R} = - \csc\theta\, \partial_{a}(\sin\theta {}^{(2)}\! W^{a}) \, . \label{calR=dW} \end{equation} Therefore, the term in (\ref{limk-k0}) which is enclosed by square brackets vanishes, and we obtain \begin{equation} \lim_{R \rightarrow \infty} \int_{B(w_{*},R)} {\rm d}^{2}x \sqrt{\sigma} N \varepsilon = \frac{2}{\kappa}\int {\rm d}\Omega{\,} \alpha(\theta,\phi) {\rm m} (w_{*},\theta,\phi) {\,} , \label{limitresult} \end{equation} for the desired limit. This result shows that the $R\to\infty$ limit of the smeared energy surface density equals the smeared mass aspect. Coupled with the findings of the next section, it follows that Geroch's ``supermomentum" is just the smeared mass aspect. This simple result does not appear to be widely known. Finally, recall that the Bondi-Sachs four-momentum components\footnote{Underlined Greek indices refer to components of the total Bondi-Sachs four-momentum.} $P_{_{BS}}^{\underline{\lambda}}$ correspond asymptotically to a pure translation. In terms of the smeared energy surface density, one obtains a pure translation for a judicious choice of lapse function on $B(w_{*},\infty)$; namely, $\alpha(\theta,\phi) = \epsilon_{\underline{\lambda}} \alpha^{\underline{\lambda}}(\theta,\phi)$, where the $\epsilon_{\underline{\lambda}}$ are constants and \cite{Goldberg} \begin{eqnarray} \alpha^{\underline{0}} & = & 1 \label{translations} \eqnum{\ref{translations}a} \\ \alpha^{\underline{1}} & = & \sin\theta\cos\phi \eqnum{\ref{translations}b} \\ \alpha^{\underline{2}} & = & \sin\theta\sin\phi \eqnum{\ref{translations}c} \\ \alpha^{\underline{3}} & = & \cos\theta \eqnum{\ref{translations}d}\, . \addtocounter{equation}{1} \end{eqnarray} Therefore, we write $\epsilon_{\underline{\lambda}} P_{_{BS}}^{\underline{\lambda}}(w_{*}) = \lim_{R \rightarrow \infty} H_{B(w_{*},R)}$ for the appropriate limiting value of $N$, and thereby obtain the Bondi-Sachs four-momentum as a Hamiltonian value. \section{Supermomentum} In this section we show that the null limit of the smeared Hamiltonian boundary value, Eq.~(4.1), is the ``supermomentum" of Geroch.\cite{Geroch} To be precise, we show that in the null limit $H_{B}$ equals Geroch's ``supermomentum" as written by Dray and Streubel.\cite{DrayStreubel} The spin-coefficient formalism is required for this analysis.\footnote{Throughout $\S$V we deal exclusively with smooth expansions in inverse powers of an {\em affine} radius, as we know of no work examining the standard spin coefficient approach to null infinity within a more general framework such as the polyhomogeneous one. The expansions we borrow from \cite{Dougan} are valid for Einstein-Maxwell theory.} Apart from a few minor notational changes we adopt the conventions of Dougan.\cite{Dougan} Geometrically, the scenario is nearly the same as the one described in the previous sections. However, we now work with a slightly different type of Bondi coordinates. Namely, $(w,r,\zeta,\bar{\zeta})$, where $r$ is an {\em affine} parameter along the null-geodesic generators of ${\cal N}$ and $\zeta = e^{{\rm i}\phi}\cot(\theta/2)$ is the stereographic coordinate. Dougan picks\footnote{Our $k_{\mu}$ and $l_{\mu}$ respectively correspond to $l_{a} = \nabla_{a} u$ and $n_{a}$ in \cite{Dougan}, where $u$ is Dougan's retarded time. The minus sign difference between our definition for $k_{\mu}$ and Dougan's definition for $l_{a}$ stems from a difference in metric-signature conventions [ours is $(-,+,+,+)$]. The convention for metric signature does not affect the spin coefficients (\ref{spinexpansions}).} $k_{\mu} = - \nabla_{\mu} w$ as the first leg of a null tetrad, which is the same normal as given in (\ref{bassae}a) if $\eta = 0$. For convenience, in this section we ignore the kinematical freedom associated with the $\eta$ parameter, setting it to zero throughout. As before, the vector field $u^{\mu} := \frac{1}{2} k^{\mu} + l^{\mu}$ (equality restricted to ${\cal N}$) defines a three-surface $\Sigma$ spanning each $B$ slice of ${\cal N}$. It follows that $u^{\mu}\partial/\partial x^{\mu} \rightarrow \partial/\partial w$ as $r \rightarrow \infty$, and hence our asymptotic slice $\Sigma_{\infty}$ again defines a pure {\sc bms} time translation. Let us first collect the essential background results from \cite{Dougan} which we will need. First, the required spin coefficients have the following asymptotic expansions:\footnote{Note the dual use of $\sigma$ as both the stem letter for the $B$ two-metric and as the spin coefficient known as the shear. We have used $\sigma$ twice in order to stick with the conventions of our references as much as possible. In all but equation (\ref{spinexpansions}), where $\sigma$ has the spin-coefficient meaning, it carries a ``$0$'' superscript denoting the asymptotic piece.} \begin{eqnarray} \rho & = & - r^{-1} - \sigma^{0} \bar{\sigma}{}^{0} r^{-3} + O(r^{-5}) \nonumber \\ \mu & = & - {\textstyle \frac{1}{2}} r^{-1} - [\Psi^{0}_{2} + \sigma^{0} \dot{\bar{\sigma}}{}^{0} + \mbox{$\partial \hspace{-2mm} /$}^{2}_{{\scriptscriptstyle 0}} \bar{\sigma}{}^{0}]r^{-2} + O(r^{-3}) \nonumber \\ \sigma & = & \sigma^{0} r^{-2} + O(r^{-4}) \label{spinexpansions} {\,} , \end{eqnarray} where $\mu$ is Dougan's $- \rho'$, the term $\Psi^{0}_{2}$ is a certain asymptotic component of the Weyl tensor, and $\sigma^{0}$ is the asymptotic piece of the shear. Like before, an over-dot denotes differentiation by $\partial/\partial w$. As fully described in \cite{Dougan}, $\ed{}_{\!\naught}$ is the standard differential operator from the compacted spin-coefficient formalism, here defined on the {\em unit} sphere. The expansion for the corresponding full operator in spacetime is \begin{equation} \mbox{$\partial \hspace{-2mm} /$} = r^{-1} \ed{}_{\!\naught} + r^{-2} [s(\bar{\ed}{}_{\!\naught}\sigma^{0}) - \sigma^{0} \bar{\ed}{}_{\!\naught}] + O(r^{-3}) {\,} , \end{equation} where $s =$ {\sc sw}$(\varphi)$, $\varphi$ being the spacetime scalar on which $\mbox{$\partial \hspace{-2mm} /$}$ acts and {\sc sw} denoting {\em spin weight}. The commutator of $\mbox{$\partial \hspace{-2mm} /$}$ and $\bar{\mbox{$\partial \hspace{-2mm} /$}}$ is \begin{equation} (\bar{\mbox{$\partial \hspace{-2mm} /$}}\mbox{$\partial \hspace{-2mm} /$} - \mbox{$\partial \hspace{-2mm} /$} \bar{\mbox{$\partial \hspace{-2mm} /$}})\varphi = {\textstyle \frac{1}{2}} s {\cal R} \varphi {\,} . \label{commutator} \end{equation} Now consider the following ansatz for the $B$ intrinsic Ricci scalar: \begin{equation} {\cal R} = 2r^{-2} + {}^{(3)}\!{\cal R} r^{-3} + O(r^{-4}) {\,} . \end{equation} If we insert this expansion into (\ref{commutator}) and expand both sides of the equation [assuming $\varphi = {}^{(0)}\!\varphi + {}^{(1)}\!\varphi r^{-1} + O(r^{-2})$ with {\sc sw}$(\varphi) = 1$], then to lowest order, namely $O(r^{-2})$, we get a trivial equality. However, equality at the next order demands that \begin{equation} {\textstyle \frac{1}{2}} {}^{(3)}\!{\cal R} = \bar{\ed}{}_{\!\naught}^{2} \sigma^{0} + \ed{}_{\!\naught}^{2} \bar{\sigma}{}^{0} {\,} . \label{result} \end{equation} This will prove to be a very important result for our purposes. Finally, Dougan gives the following expansion for the $B$ volume element: \begin{equation} {\rm d}^{2}x\sqrt{\sigma} = {\rm d}\Omega {\,} r^{2} (1 - \sigma^{0} \bar{\sigma}^{0} r^{-2}) + O(r^{-2}) {\, } . \end{equation} (Here $\sigma$ is the determinant of the $B$ metric and $\sigma^0$ is the asymptotic piece of the shear.) We now consider the spin-coefficient expression for the smeared energy surface density introduced in $\S$IV. Again with $k = 2\mu + \rho$, in the present notation we find \begin{equation} k = - 2r^{-1} - 2[\Psi^{0}_{2} + \sigma^{0} \dot{\bar{\sigma}}{}^{0} + \mbox{$\partial \hspace{-2mm} /$}^{2}_{{\scriptscriptstyle 0}} \bar{\sigma}{}^{0}]r^{-2} + O(r^{-3}) {\,} . \end{equation} Moreover, by an argument identical to the one found in the last paragraph of the Appendix (although here with the affine radius $r$ rather than the areal radius $R$), we know that the result (\ref{result}) determines \begin{equation} k |^{\scriptscriptstyle {\rm ref}} = - 2 r^{-1} -(\bar{\ed}{}_{\!\naught}^{2} \sigma^{0} + \ed{}_{\!\naught}^{2} \bar{\sigma}{}^{0}) r^{-2} + O(r^{-3}) \end{equation} as the appropriate asymptotic expansion for the reference term. Therefore, ($\kappa$ times) the full quasilocal energy surface density is \begin{equation} \kappa\varepsilon = - 2[\Psi^{0}_{2} + \sigma^{0} \dot{\bar{\sigma}}{}^{0} + {\textstyle \frac{1}{2}}\ed{}_{\!\naught}^{2} \bar{\sigma}{}^{0} - {\textstyle \frac{1}{2}}\bar{\ed}{}_{\!\naught}^{2} \sigma^{0} ]r^{-2} + O(r^{-3}) {\,} . \end{equation} At this point we consider again a smearing function $N$, with appropriate fall-off behavior and limit $\alpha(\zeta,\bar{\zeta}) = \lim_{r \rightarrow \infty} N(r,\zeta,\bar{\zeta})$. Using the results amassed up to now, one computes that the limit of the smeared energy surface density is \begin{equation} \lim_{r \rightarrow \infty} \int_{B(w_{*}, r)} {\rm d}^{2}x \sqrt{\sigma} N\varepsilon = - \frac{2}{\kappa} \int {\rm d}\Omega{\,}\alpha\left. [\Psi^{0}_{2} + \sigma^{0} \dot{\bar{\sigma}}{}^{0} + {\textstyle \frac{1}{2}}\ed{}_{\!\naught}^{2} \bar{\sigma}{}^{0} - {\textstyle \frac{1}{2}}\bar{\ed}{}_{\!\naught}^{2} \sigma^{0}]\right|_{w = w_{*}} {\,} . \label{superE} \end{equation} The right-hand side of this equation is the ``supermomentum'' of Geroch as written by Dray and Streubel [see equation (A1.12) of \cite{DrayStreubel} and set their $b = 0$ for a Bondi frame as we have here]; and the ``supermomentum'' is known to be the ``charge integral" associated with the Ashtekar-Streubel flux\cite{AshtekarStreubel} of gravitational radiation at ${\cal I}^{+}$ (in the restricted case when the flux is associated with a supertranslation).\cite{Shaw} Dray and Streubel have discussed the importance of the particular factors of $\frac{1}{2}$ which multiply the last two terms within the square brackets on the right-hand side of equation (\ref{superE}). It is evident from our approach that the origin of these $\frac{1}{2}$ factors stems from the flat-space reference of the quasilocal energy (flat-space being the correct reference in the present context). When $\alpha$ determines a pure {\sc bms} translation, the last two terms in the integrand integrate to zero. For instance, setting $\alpha = 1$, one finds that the strict energy \begin{equation} E = \int_{B(w_{*},\infty)} {\rm d}^{2}x\sqrt{\sigma} \varepsilon = - \frac{2}{\kappa} \int {\rm d}\Omega \left. [\Psi^{0}_{2} + \sigma^{0} \dot{\bar{\sigma}}{}^{0}]\right|_{w = w_{*}} {\,} \end{equation} is the standard spin-coefficient expression for the Bondi-Sachs mass $M_{_{BS}}(w_{*})$.\cite{Dougan,ExtonNewmanPenrose} \section*{Acknowledgments} We thank H. Balasin, P. T. Chru\'{s}ciel, T. Dray, and N. \'{O} Murchadha for helpful discussions and correspondence. We acknowledge support from National Science Foundation grant 94-13207. S.\ R.\ Lau has been chiefly supported by the ``Fonds zur F\"{o}r\-der\-ung der wis\-sen\-schaft\-lich\-en For\-schung'' in Austria (FWF project P 10.221-PHY and Lise Meitner Fellowship M-00182-PHY).
proofpile-arXiv_065-464
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
proofpile-arXiv_065-465
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} Blazars are variable, polarized, flat spectrum extragalactic radio sources with a non-thermal continuum extending to $\gamma$-ray energies ({\it e.g.,\ } Urry \& Padovani 1995). The radio emission from blazars is collimated into narrow beams composed of many individual knots and an optically thick core (Kellerman \& Pauliny-T\"oth 1981). In virtually all blazars, the radio knots appear to separate from the core at speeds greater than the speed of light (Urry \& Padovani 1995) and this superluminal motion is strong evidence that blazars are relativistic jets of magnetized plasma viewed along the jet axis (Blandford \& K\"onigl 1979). Although the jet model also accounts for such diverse blazar properties as the flat radio spectrum (Kellerman \& Pauliny-T\"oth 1981), the short variability time scales at all energies ({\it e.g.,\ } Quirrenbach, {\it et~al.\ } 1991, Maraschi, {\it et~al.\ } 1992), the small synchrotron self-Compton X-ray fluxes (Marscher 1987, Ghisellini, {\it et~al.\ } 1992) and the large $\gamma$-ray fluxes (Maraschi, {\it et~al.\ } 1992), the radiative processes which produce the blazar continuum have not been identified. In this paper, we calculate the relative amplitudes of variations in the radio and X-ray fluxes for one of the most popular models of the blazar continuum, synchrotron self-Compton scattering in a relativistic outflow (Jones, O'Dell \& Stein 1974, Marscher 1977, K\"onigl 1981, Ghisellini, {\it et~al.\ } 1985), and we compare the results to observations of correlated variations on the radio and X-ray flux from 3C279 (Grandi, etal 1995: G95). Blazar spectra are nearly featureless and a large number of models, which make very different assumptions for the relevant radiation processes, agree qualitatively with snapshots of the radio to $\gamma$-ray spectrum (Maraschi, {\it et~al.\ } 1992, Maraschi, {\it et~al.\ } 1994, Hartmann {\it et~al.\ } 1996). At low energies, the flat radio spectrum of compact radio cores can be explained by inhomogeneous synchrotron emission (Marscher 1977, K\"onigl 1981, Ghisellini, {\it et~al.\ } 1985) or by a homogeneous core that is optically thick to induced Compton scattering (Sincell \& Krolik 1994: SK94). The situation at high energies is even more complicated. The X- and $\gamma$-ray emission may be synchrotron emission by high energy electrons (K\"onigl 1981, Ghisellini, {\it et~al.\ } 1985), radiation from a pair cascade (Blandford \& Levinson 1995) or low frequency photons which are inverse Compton scattered by relativistic electrons in the jet. In the last case, the source of the low energy photons could be synchrotron radiation from the jet (the synchrotron self-Compton model, Maraschi {\it et~al.\ } 1992, Maraschi, {\it et~al.\ } 1994, Bloom \& Marscher 1996), UV radiation from a disk (Dermer, Schlickeiser \& Mastichiadis 1992), or some diffuse source of radiation surrounding the jet (Sikora, Begelman \& Rees 1994, Ghisellini \& Madau 1996). We will consider only synchrotron self-Compton scattering in this paper and ignore other sources for the high energy emission. The synchrotron self-Compton model for the continuum emission assumes that a single population of relativistic electrons radiates synchrotron photons and subsequently scatters a fraction of these to higher energies (Jones, {\it et~al.\ } 1974). Fluctuations in either the electron density, the magnetic field strength or the Doppler factor of the emission region will affect the synchrotron and self-Compton fluxes instantaneously. Therefore, variations in the low and high energy fluxes which are uncorrelated, or have significant time delays, cannot be the result of synchrotron self-Compton scattering. The relative amplitudes of the variations in the synchrotron and self-Compton fluxes depends upon which physical parameters change. For example, the fractional change in the high energy flux is twice as large as the fractional change in the optically thin synchrotron flux when the electron density varies, whereas they are equal if the magnetic field or the Doppler factor changes (SK94). The amplitude of the variations in the synchrotron and self-Compton fluxes also depends upon the relativistic electron distribution. The electron spectrum will flatten with increasing radiation intensity because both the synchrotron and inverse Compton cooling rates increase with increasing radiation energy density. However, these cooling processes are both too slow to have any effect upon the electron distribution in the parsec scale jet (SK94). Therefore, we neglect the evolution of the electron spectrum in the calculations presented in this paper. SK94 demonstrated that induced Compton scattering can reduce the amplitudes of variations in the synchrotron radiation when the brightness temperature of the source $T_B \mathrel{\raise.3ex\hbox{$<$\kern-.75em\lower1ex\hbox{$\sim$}}} 2 \times 10^{11} \mbox{K}$. The amplitudes of variations in the self-Compton X-ray flux are unaffected by induced Compton scattering because the X-ray flux is dominated by photons scattered from the high-frequency, low $T_B$, end of the synchrotron spectrum, where the stimulated scattering optical depth is small. Thus, the relative amplitudes of the variations in the synchrotron flux at the self-absorption turnover frequency and the self-Compton flux at 1 keV are reduced from $\sim 0.4$ to $\sim 0.2$ when the induced Compton scattering optical depth is large (SK94). The continuum emission from 3C279, one of the most intensively monitored blazars, varies coherently over its entire spectrum (Maraschi, {\it et~al.\ } 1992, Maraschi {\it et~al.\ } 1994, Hartmann, {\it et~al.\ } 1996). Recently, G95 used historical light curves to show that the radio-mm and 1 keV X-ray fluxes from 3C279 are strongly correlated. The maximum time resolution of the 3C279 light curves was $\sim 70 \mbox{d}$ and the absence of any detectable time delay implies that the radio-mm and X-ray fluxes are from physically related regions separated by $\mathrel{\raise.3ex\hbox{$<$\kern-.75em\lower1ex\hbox{$\sim$}}} 0.06 \mbox{pc}$. While this is consistent with the assumptions of the synchrotron self-Compton theory, the relative amplitudes of the radio and X-ray variations are smaller than predicted (Sincell 1996). In this paper, we extend previous calculations (SK94, Sincell 1996) and compute the relative amplitudes of variations in the synchrotron self-Compton flux as a function of the frequency of the synchrotron emission. We first define the flux variability ratio (\S \ref{sec: flux variability ratio}) and describe how it is calculated. This ratio is computed for three simple types of variations in \S\ref{sec: results} and the implications for 3C279 are discussed in \S \ref{sec: 3C279}. We conclude in \S \ref{sec: conclusions}. \section{\bf The Flux Variability Ratio} \label{sec: flux variability ratio} The time variability of the source spectrum is approximated as a sequence of steady state spectra. We have used the code developed in SK94 to compute the steady-state synchrotron spectrum and self-Compton X-ray flux from a homogeneous sphere containing an isotropic power-law distribution of relativistic electrons \begin{equation} \label{eq: electron distribution} {\partial n_e \over \partial \gamma} = n_o \gamma^{-p} \end{equation} for $\gamma \geq \sqrt{2}$. We assume $p=2.5$ in all the simulations and the normalization ($n_o$) is fixed by the assumed $\tau_T$. This code incorporates synchrotron absorption and emission, inverse Compton scattering and induced Compton scattering by the relativistic electrons. Stimulated scattering becomes the dominant source of radio opacity when $T_B \mathrel{\raise.3ex\hbox{$>$\kern-.75em\lower1ex\hbox{$\sim$}}} 2 \times 10^{11} \mbox{K}$ (SK94) and must be included when calculating the spectra of compact radio sources. Self-absorption reemerges as the dominant opacity source at low frequencies and inverse Compton scattering of synchrotron photons by electrons with $\gamma \mathrel{\raise.3ex\hbox{$<$\kern-.75em\lower1ex\hbox{$\sim$}}} 10$ contributes to the radio flux above the self-absorption turnover. The flux variability ratio \begin{equation} R_m(\tau_T,B,\nu) = {\partial \log S_r(\tau_T,B,\nu) \over \partial \log S_x(\tau_T,B)}|_m, \end{equation} is the ratio of the fractional change in the synchrotron flux at $\nu$ ($S_r$) and the self-Compton X-ray flux ($S_x$) at 1 keV caused by a fluctuation in the physical parameter $m$. In this paper we investigate three different variations: the Thomson depth ($\tau_T$) of the source varies at fixed magnetic field strength ($m=\tau$), the magnetic field strength ($B$) varies at fixed $\tau_T$ ($m=B$) and equal fractional changes in $\tau_T$ and $B$ $(m=S)$. The third case approximates the passage of a strong shock through the plasma, assuming that the field is tangled ({\it e.g.,\ } Marscher \& Gear 1985). We also assume that $\tau_T$ and $B$ are uniform throughout the source. Simple analytic calculations of $R_m$ are possible for synchrotron self-Compton scattering (SK94), but $R_m$ must be calculated numerically when the induced Compton scattering opacity is large (SK94). We compute the variability ratio using the approximate formula \begin{equation} R_m(\nu) \simeq {S_r(m,\nu) - S_r(m+\Delta m,\nu) \over S_r(m,\nu) + S_r(m+\Delta m,\nu)} \cdot {S_x(m) + S_x(m+\Delta m) \over S_x(m) - S_x(m+\Delta m)}, \end{equation} and the radio spectra and X-ray fluxes from two models with closely spaced values of the parameter $m$. In this paper we choose $\Delta m/m =0.1$ but the results are fairly insensitive to $\Delta m$, even when $\Delta m/m \mathrel{\raise.3ex\hbox{$>$\kern-.75em\lower1ex\hbox{$\sim$}}} 1$ (see figures in Sincell 1996). \section{Results} \label{sec: results} We have calculated $R_{\tau}$, $R_{B}$ and $R_S$ for $0.01 \leq \tau_T \leq 3.0$ and a range of $B$. In figs. \ref{fig: tau variability}, \ref{fig: b variability} and \ref{fig: shock variability} we plot $R_m$ as a function of $\nu / \nu_o$, where $\nu_o$ is the synchrotron self-absorption turnover frequency. When the effects of induced and inverse Compton scattering on the radio spectrum are neglected, $R_m(\nu/\nu_o)$ is independent of the unperturbed values of $\tau_T$ and $B$. Including Compton scattering in the computation of the radio spectrum introduces a strong dependence upon $\tau_T$ but $R_m(\nu/\nu_o)$ remains nearly independent of $B$ because the induced Compton scattering optical depth at $\nu_o$ is $\propto T_B \tau_T \propto B^{-1/(p+4)} \tau_T^{(p+5)/(p+4)}$ (SK94). The code was used to verify that $R_m$ changed by less than a few percent when $B$ increased by two orders of magnitude. Therefore, we present $R_m$ for $B=10^{-5} \mbox{G}$ and use the relation ({\it e.g.,\ } SK94) \begin{equation} \label{eq: nuo scaling} \nu_o \propto \left( {\delta \over 1+z} \right) B^{(p+2)/(p+4)} \end{equation} to scale $R_m$ to any desired field strength, Doppler boost ($\delta$) or redshift ($z$). Although $R_m(\nu/\nu_o)$ is independent of the unperturbed values of $\tau_T$ and $B$ when the effects of Compton scattering on the radio spectrum are neglected, it does depend on which parameters vary. At optically thin frequencies, $\nu \gg \nu_o$, $R_{\tau} \simeq 0.5$, $R_{B} \simeq 1.0$ and $R_{S} \simeq 0.7$, independent of frequency (figs. \ref{fig: tau variability}, \ref{fig: b variability} and \ref{fig: shock variability}). These numerical values are in good agreement with the analytic results in SK94. Synchrotron self-absorption reduces $R_m$ at $\nu \mathrel{\raise.3ex\hbox{$<$\kern-.75em\lower1ex\hbox{$\sim$}}} \nu_o$. This is because any change in the physical parameters which increases the synchrotron emissivity results in a compensating increase in the opacity and decrease in the photospheric depth. Thus, the net flux at optically thick frequencies is less sensitive to changes in the source parameters. The ratio of the self-absorption opacity to the emissivity increases rapidly as $\nu/\nu_o$ decreases ({\it e.g.,\ } SK94) and $R_{\tau}$ approaches zero at $\nu \mathrel{\raise.3ex\hbox{$<$\kern-.75em\lower1ex\hbox{$\sim$}}} \nu_o / 2$. Increasing $B$ also increases $\nu_o$ (eq. \ref{eq: nuo scaling}) and the increase in the self-absorption opacity at fixed $\nu$ overwhelms the increase in the emissivity when $\nu \ll \nu_o$. The resulting decrease in the synchrotron flux at $\nu$ appears as negative values of $R_{B}$ and $R_{S}$ (figs. \ref{fig: b variability} and \ref{fig: shock variability}). Negative values of $R_m$ correspond to an anti-correlation of the synchrotron and self-Compton fluxes. Compton scattering changes the frequency dependence of $R_m$ and introduces a dependence on $\tau_T$ (figs. \ref{fig: tau variability}, \ref{fig: b variability} and \ref{fig: shock variability}). Induced Compton scattering reduces $R_{\tau}$ over more than a decade in frequency when $\tau_T \mathrel{\raise.3ex\hbox{$>$\kern-.75em\lower1ex\hbox{$\sim$}}} 0.1$ (fig. \ref{fig: tau variability}), relative to the synchrotron self-Compton scattering model without stimulated scattering. $R_{\tau}$ at $\nu \simeq \nu_o$ is reduced by almost an order of magnitude when $\tau_T \mathrel{\raise.3ex\hbox{$>$\kern-.75em\lower1ex\hbox{$\sim$}}} 1$. The stimulated scattering opacity at low frequencies and the contribution of inverse Compton scattered photons at higher frequencies results in a monotonic increase in $R_{\tau}$ from $0.01$ at $\nu \mathrel{\raise.3ex\hbox{$<$\kern-.75em\lower1ex\hbox{$\sim$}}} \nu_o$ to $0.5$ at $\nu \sim 100 \nu_o$ (fig. \ref{fig: tau variability}). Even though synchrotron self-absorption is the dominant source of opacity, stimulated scattering increases the photon occupation numbers at low frequencies. This increases the synchrotron flux at $\nu \ll \nu_o$ and the anti-correlation of the synchrotron and self-Compton fluxes caused by variations in $B$ disappears when $\tau_T \mathrel{\raise.3ex\hbox{$>$\kern-.75em\lower1ex\hbox{$\sim$}}} 1$. Variations in the magnetic field strength introduce local extrema into $R_{B,S}$ when the stimulated scattering optical depth is large. The largest contribution to the induced Compton scattering opacity at $\nu \mathrel{\raise.3ex\hbox{$<$\kern-.75em\lower1ex\hbox{$\sim$}}} \nu_o$ is from electrons with $\gamma_{*} = {1 \over 2} \left( {\nu_m \over \nu} \right)^{1/2}$ where $\nu_m \mathrel{\raise.3ex\hbox{$>$\kern-.75em\lower1ex\hbox{$\sim$}}} \nu_o$ is the peak of the spectrum (SK94). The low energy cutoff $\gamma = \sqrt{2}$ reduces the stimulated scattering opacity at frequencies $ \nu_m/8 \mathrel{\raise.3ex\hbox{$<$\kern-.75em\lower1ex\hbox{$\sim$}}} \nu \mathrel{\raise.3ex\hbox{$<$\kern-.75em\lower1ex\hbox{$\sim$}}} 8\nu_m$ because $\gamma_* < \sqrt{2}$ and there are no electrons which couple $\nu$ to $\nu_m$. When synchrotron self-absorption is the dominant source of opacity, the optical depth of the plasma decreases with frequency and $R$ increases with frequency. A local maximum in $R$ appears at the frequency where the stimulated scattering opacity is approximately equal to the synchrotron opacity. At higher frequencies, induced Compton scattering limits the variations in the synchrotron flux and reduces $R$. The local minimum in $R$ occurs at $\nu \sim \nu_m/8$ where the stimulated scattering opacity reaches a maximum. These local extrema are not as prominent in $R_{\tau}$ because variations in $\tau_T$ increase the optical depth at all frequencies. However, the kink in the $\tau_T = 0.1$ curve of $R_{\tau}$ (fig. \ref{fig: tau variability}) is also due to this effect. \section{\bf 3C279} \label{sec: 3C279} G95 used historical light curves of 3C279 to show that variations in the radio-mm and X-ray fluxes are strongly correlated with a time delay of $\mathrel{\raise.3ex\hbox{$<$\kern-.75em\lower1ex\hbox{$\sim$}}} 70$~days. They also calculated the logarhithmic dispersion, or variability amplitude ($v(\nu)$), of the available measurements and found that $v(\nu)$ increases systematically with frequency. However, these estimates of $v(\nu)$ are uncertain because many emission components contribute to the observed flux at a given frequency ({\it e.g.,\ } Unwin, {\it et~al.\ } 1989). These components may vary independently and the G95 data lacks the spatial resolution necessary to reliably subtract the non-variable background. Long-term monitoring at higher resolution (VLBA) is necessary to remove this source of uncertainty. In the remainder of this paper we will assume that the variable component dominates the observed flux, but it should be remembered that a significant amount of non-variable flux at $\nu$ will reduce $v(\nu)$ below the value expected for the variable component alone. Both the strong correlation of the radio and X-ray fluxes and the absence of a detectable time delay between variations in the two bands are consistent with the assumptions of the synchrotron self-Compton model for the continuum emission. We used the variability amplitudes calculated by G95 to estimate \begin{equation} R \simeq {v(\nu) \over v(1 \mbox{keV})} \end{equation} at four frequencies in the range $14.5 \mbox{GHz} < \nu < 230 \mbox{GHz}$. The estimated $R$ for the epochs 1988-1991.4 and 1991.4-1993.2 are plotted in fig. \ref{fig: observed variability}. The model $R_{\tau}$ for $\tau_T = 1.0$, $B=10^{-3} \mbox{G}$ and $\delta = 20$ is displayed on the same figure. The relative amplitudes of the variations in the radio and X-ray fluxes from 3C279 are consistent with the synchrotron self-Compton model if $\tau_T$ varies in a fixed magnetic field and induced Compton scattering is the dominant source of radio opacity. It is immediately apparent from fig. \ref{fig: observed variability} that the magnitudes of both $R_{B}$ and $R_{S}$ are too large to fit the observed values of $R$ for 3C279. In addition, neither the local extrema in $R$ or the anti-correlation of the synchrotron and self-Compton fluxes are observed. We also find that the increase in $R$ with frequency is much slower than expected for the synchrotron self-Compton model without induced Compton scattering (fig. \ref{fig: observed variability}). However, both the magnitude and the frequency dependence of $R$ are consistent with $R_{\tau}$ when $\tau_T \mathrel{\raise.3ex\hbox{$>$\kern-.75em\lower1ex\hbox{$\sim$}}} 1$. This implies that $\tau_T \sim 1$ in the synchrotron self-Compton emission region. The magnetic field strength and Doppler factor cannot be calculated independently (eq. \ref{eq: nuo scaling}), but the values we have assumed ($B = 10^{-3} \mbox{G}$ and $\delta = 20$) are consistent with other estimates of the physical parameters for 3C279 (Ghisellini, etal 1985, Maraschi, {\it et~al.\ } 1992, SK94). Larger magnetic fields require smaller Doppler factors and vice versa. We can set a lower limit on the electron density using the variability time scale and the requirement $\tau_T \sim 1$. The linear dimension of the emission region $l \mathrel{\raise.3ex\hbox{$<$\kern-.75em\lower1ex\hbox{$\sim$}}} l_{max} = 1.8 \times 10^{17} \delta \mbox{cm}$ if the observed flux varies on a time scale of $\mathrel{\raise.3ex\hbox{$<$\kern-.75em\lower1ex\hbox{$\sim$}}} 70$ days. The stimulated scattering optical depth of the plasma will be large enough to reduce the variability amplitudes if the electron density $n_e \mathrel{\raise.3ex\hbox{$>$\kern-.75em\lower1ex\hbox{$\sim$}}} 8 \times 10^6 \delta^{-1} \mbox{ ${\rm cm^{-3}}$}$. The total particle energy density of the distribution in eq. \ref{eq: electron distribution} is dominated by the rest mass energy so $U_e \sim n_e m_p c^2 \sim 7 \delta^{-1} \left( {m_p \over m_e} \right) \mbox{ergs ${\rm cm^{-3}}$}$, where $m_{p,e}$ are the masses of the positively charged particle and an electron, respectively. The magnetic energy density $U_B \ll U_e$ and the plasma is far from equipartition. A strong shock passing through the synchrotron emission region amplifies both the electron density and the magnetic field strength if the magnetic field is either tangled or aligned parallel to the shock front ({\it e.g.,\ } Marscher \& Gear 1985). Our results for 3C279 indicate that the variations in the flux are due to fluctuations in the electron density alone. Thus, we conclude that if the variations are caused by a shock the magnetic field must be aligned perpendicular to the shock front. An alternative possibility is that the observed variations are due to fluctuations in the local electron density caused by a change in the particle injection rate. \section{\bf Conclusions} \label{sec: conclusions} We have calculated $R_m(\nu)$, the relative amplitude of variations in the synchrotron flux at $\nu$ and the self-Compton X-ray flux at 1 keV, for a homogeneous sphere of relativistic electrons orbiting in a tangled magnetic field. The index $m$ refers to the physical quantity which is assumed to vary and in this paper we investigate three cases: variations in $\tau_T$ at fixed $B$, variations in $B$ at fixed $\tau_T$ and equal fractional changes in both quantities. The last case approximates the passage of a strong shock through the plasma ({\it e.g.,\ } Marscher \& Gear 1985). We find $R_m$ to be useful for two reasons. First, the frequency dependence of $R_m$ is determined by the optical depth of the plasma. Second, $R$ may be estimated directly from observations of correlated radio and X-ray variability ({\it e.g.,\ } G95). If synchrotron self-absorption is the dominant source of opacity, the frequency dependence of $R_m$ is determined by $\nu_o$, the self-absorption turnover frequency, and the physical parameter which is assumed to vary. We find that $R_m$ is constant at all optically thin frequencies ($\nu \gg \nu_o$) and, for our assumed electron distribution (eq. \ref{eq: electron distribution}), $R_{\tau} \simeq 0.5$, $R_B \simeq 1.0$ and $R_S \simeq 0.7$. Self-absorption causes all the $R_m$ to drop sharply at $\nu \mathrel{\raise.3ex\hbox{$<$\kern-.75em\lower1ex\hbox{$\sim$}}} \nu_o$ and both $R_{B,S}$ become negative at $\nu \ll \nu_o$. Induced Compton scattering reduces $R_m$ over more than a decade in frequency, relative to the synchrotron self-Compton model without stimulated scattering, when $\tau_T \mathrel{\raise.3ex\hbox{$>$\kern-.75em\lower1ex\hbox{$\sim$}}} 0.1$. Increasing $\tau_T$ reduces $R_{\tau}$ at frequencies near $\nu_o$ and $R_{\tau}$ can be an order of magnitude smaller than the self-absorbed value when $\tau_T \mathrel{\raise.3ex\hbox{$>$\kern-.75em\lower1ex\hbox{$\sim$}}} 1$. We find that $R_{\tau}$ increases monotonically from low to high frequencies, but a slight change in the stimulated scattering opacity at $\nu \sim \nu_m / 8$ causes local extrema in $R_{B,S}$. Variations in the Thomson depth of a homogeneous source of synchrotron self-Compton radiation reproduces the relative amplitudes of the correlated radio and X-ray flux variations in 3C279 (G95) if $\tau_T \sim 1$ and the emission region is optically thick to induced Compton scattering. Although $B$ and $\delta$ cannot be independently constrained (eq. \ref{eq: nuo scaling}), the observed $R$ is consistent with $B \sim 10^{-3} \mbox{G}$ and $\delta \sim 20$. If we assume that the maximum linear dimension of the emission region is $l_{max} \sim 0.06 \delta$ pc, as implied by the variability time scale, $\tau_T \sim 1$ requires that the electron density be $n_e \mathrel{\raise.3ex\hbox{$>$\kern-.75em\lower1ex\hbox{$\sim$}}} 8 \times 10^6 \delta^{-1}$~cm$^{-3}$. In this case, the particle energy density is much larger than the magnetic field energy density. Variations in the magnetic field strength result in values of $R$ which are larger than observed. If the observed fluctuations are due to the passage of a shock, the magnetic field must be oriented perpendicular to the shock front. Variations in the local particle injection rate could change the electron density without necessarily changing the field strength. Finally, it has been argued that stimulated scattering cannot be important in blazars with optically thick spectral indices of $\alpha = -5/2$ (Litchfield, {\it et~al.\ } 1995). This argument is erroneous because synchrotron self-absorption {\it always} becomes the dominant source of opacity, and $\alpha = -5/2$, at low enough frequencies (SK94). However, this points out the ambiguities inherent in attempting to determine the radio opacity using spectral measurements alone. Additional multi-wavelength variability studies ({\it e.g.,\ } G95) are clearly necessary to determine the relevant radiative processes in blazars. \acknowledgements We thank the referee, Laura Maraschi, for helpful comments and Stephen Hardy for pointing out the error in the kinetic equation. Support for this work was provided by NASA grants NAGW-1583 and NAG 5-2925, and NSF grant AST 93-15133. The simulations were performed at the Pittsburgh Supercomputing Center (grant AST 960002P).
proofpile-arXiv_065-466
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} We propose a new search for the time reversal violating polarization of the muon normal to the plane of the $K^+ \to \mu^+ \pi^0 \nu$ decay~\cite{proposal}. The term $\vec{\sigma_\mu} \cdot (\vec{P_\pi} \times \vec{P_\mu})$, which is proportional to the projection of the muon polarization out of the decay plane, changes sign upon time reversal; therefore a finite expectation value for this quantity indicates a violation of time reversal invariance. Moreover, since the Standard Model prediction for such polarization is zero, and there is no final state interaction, the observation of T-violation in the $K_{\mu 3}$ decay is a discovery of T-violation beyond the Standard Model. Through the CPT theorem we know that T-invariance is intimately related to CP-invariance. Although the only observed CP-violation in the neutral kaon system can be attributed to the complex phase in the CKM matrix within the Standard Model the true nature of CP-violation is far from being revealed by the current experimental data. It is now accepted that the baryon asymmetry of the universe requires a source of CP violation stronger than that embodied in the CKM matrix. Models of non-standard CP violation that produce the baryon asymmetry could also produce effects observable in the transverse polarization. Because of the very high sensitivity of the experiment the possibility of discovering unexpected new physics should not be underestimated. The best previous experimental limits were obtained over 15 years ago with 4 GeV charged kaons~\cite{campbell} at the AGS, yielding a result of $P^T_\mu = 0.0031 \pm 0.0053$. The high intensity kaon beams available now at the AGS makes it possible to improve the limit on the polarization by more than an order of magnitude. \section{Detector Design} \begin{figure} \psfig{figure=pict1.eps,height=2.8in,width=4.5in,angle=90} \caption{Schematic of the detector. A typical $K^+\rightarrow \mu^+\pi^0\nu$ events is superimposed.} \label{pict1} \end{figure} The experiment will be performed with 2 GeV/c electrostatically seperated charged kaons decaying in flight. The beam intensity will be $2\times 10^7 K^+$'s/spill with $3\times 10^{13}$ protons on target every 3.6 sec. Figure \ref{pict1} shows the plan view of the experiment. The basic workings of the experiment are the same as the experiment in Reference 2. The detailed design is, however, optimized for a high intensity 2 GeV beam. The cylindrically symmetric detector is centered on the kaon beam. The $K^+_{\mu 3}$ ~decays of interest occur in the decay tank; the photons from the $\pi^0$ decay are detected in the calorimeter; the muon stops in the polarimeter. The decay of the stopped muon is detected in the polarimeter by wire chambers, which are arranged radially with graphite wedges that serve as absorber medium. The hit pattern in the polarimeter identifies the muon stop as well as positron direction relative to the muon stop. By selecting events with $\pi^0$ moving along the beam direction and muon moving perpendicular to the beam direction in the $K^+$ center of mass frame, the decay plane coincides with the radial wedges. A non-zero transverse muon polarization causes an asymmetry between the number of muons that decay clockwise versus the number counter-clockwise. To reduce systematic errors, a weak solenoidal magnetic field along the beam direction (70 gauss or an precessing period of $\sim 1 \mu s$) with polarity reversal every spill is applied to the polarimeter. The initial muon transverse polarization causes a small shift in the phase of the sinusoidal oscillation in the measured asymmetry. The difference in the asymmetry for the two polarities is proportional to the muon polarization in the decay plane, while the sum is proportional to the muon polarization normal to the decay plane. Compared to the previous experiment, this experiment has much better background rejection and event reconstruction. The separated $K^+$ beam should greatly reduce the accidental rate. The polarimeter is fine segmented and the analyzing power is higher. The positron signature is defined by the coincidence of signals in a pair of neighboring wedges. The larger calorimeter makes it possible to reconstruct the $\pi^0$ momentum. Together with the muon trajectory, the event can be fully reconstructed. The detector acceptance and background rejection is optimized using GEANT simulation. We expect to obtain 550 events/spill, with up to 20\% background. With an analyzing power of over 30\%, we expect to reach the statistical sensitivity of $\delta P_t = 1.3\times 10^{-4}$. At such a high statistical accuracy, much care has to be taken in reducing systematic errors. We have studied various effects that may give a false signal, such as misalignment within the polarimeter, misalignment among and detector components and the beam, asymmetry in or caused by the precessing field and inefficiencies. We believe that these errors can be made acceptably small by proper construction techniques and by using symmetries of the apparatus to internally cancel the systematic errors. In addition, we will use the T-conserving component of the muon polarization to calibrate the detector analyzing power, and samples of muon stops with no known transverse polarization, such as muons from $K_{\mu 2}$ decays, to detect any systematic bias. The proposal has been submitted to the Laboratory in Aug 1996. If approved and funded, we would like to have the first engineering run in 1998. The physics data taking will take about 2000 hours of running time.
proofpile-arXiv_065-467
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
proofpile-arXiv_065-468
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section*{Table of Contents} \noindent \S0 - INTRODUCTION \medskip \noindent \S1 - THEIR PROTOTYPE IS ${\frak K}^{tr}_{\langle\lambda_n:n< \omega \rangle}$, NOT ${\frak K}^{tr}_\lambda$! \medskip \noindent \S2 - ON STRUCTURES LIKE $(\prod\limits_n \lambda_n,E_m)_{m<\omega}$, $\eta E_m\nu =:\eta(m) = \nu(m)$ \medskip \noindent \S3 - REDUCED TORSION FREE GROUPS; NON-EXISTENCE OF UNIVERSALS \medskip \noindent \S4 - BELOW THE CONTINUUM THERE MAY BE UNIVERSAL STRUCTURES \medskip \noindent \S5 - BACK TO ${\frak K}^{rs(p)}$, REAL NON-EXISTENCE RESULTS \medskip \noindent \S6 - IMPLICATIONS BETWEEN THE EXISTENCE OF UNIVERSALS \medskip \noindent \S7 - NON-EXISTENCE OF UNIVERSALS FOR TREES WITH SMALL DENSITY \medskip \noindent \S8 - UNIVERSALS IN SINGULAR CARDINALS \medskip \noindent \S9 - METRIC SPACES AND IMPLICATIONS \medskip \noindent \S10 - ON MODULES \medskip \noindent \S11 - OPEN PROBLEMS \medskip \noindent REFERENCES \eject \section{Introduction} {\bf Context.}\hspace{0.15in} In this paper, model theoretic notions (like superstable, elementary classes) appear in the introduction but not in the paper itself (so the reader does not need to know them). Only naive set theory and basic facts on Abelian groups (all in \cite{Fu}) are necessary for understanding the paper. The basic definitions are reviewed at the end of the introduction. On the history of the problem of the existence of universal members, see Kojman, Shelah \cite{KjSh:409}; for more direct predecessors see Kojman, Shelah \cite{KjSh:447}, \cite{KjSh:455} and \cite{Sh:456}, but we do not rely on them. For other advances see \cite{Sh:457}, \cite{Sh:500} and D\v{z}amonja, Shelah \cite{DjSh:614}. Lately \cite{Sh:622} continue this paper. \medskip A class ${\frak K}$ is a class of structures with an embeddability notion. If not said otherwise, an embedding, is a one to one function preserving atomic relations and their negations. If ${\frak K}$ is a class and $\lambda$ is a cardinal, then ${\frak K}_\lambda$ stands for the collection of all members of ${\frak K}$ of cardinality $\lambda$.\\ We similarly define ${\frak K}_{\leq\lambda}$. A member $M$ of ${\frak K}_\lambda$ is universal, if every $N\in {\frak K}_{\le \lambda}$, embeds into $M$. An example is $M=:\bigoplus\limits_\lambda {\Bbb Q}$, which is universal in ${\frak K}_\lambda$ if ${\frak K}$ is the class of all torsion-free Abelian groups, under usual embeddings. We give some motivation to the present paper by a short review of the above references. The general thesis in these papers, as well as the present one is: \begin{Thesis} \label{0.1} General Abelian groups and trees with $\omega+1$ levels behave in universality theorems like stable non-superstable theories. \end{Thesis} The simplest example of such a class is the class ${\frak K}^{tr} =:$ trees $T$ with $(\omega+1)$-levels, i.e. $T\subseteq {}^{\omega\ge}\alpha$ for some $\alpha$, with the relations $\eta E^0_n\nu =: \eta\restriction n=\nu \restriction n\ \&\ \lg(\eta)\geq n$. For ${\frak K}^{tr}$ we know that $\mu^+<\lambda={\rm cf}(\lambda) <\mu^{\aleph_0}$ implies there is no universal for ${\frak K}^{tr}_\lambda$ (by \cite{KjSh:447}). Classes as ${\frak K}^{rtf}$ (defined in the title), or ${\frak K}^{rs(p)}$ (reduced separable Abelian $p$-groups) are similar (though they are not elementary classes) when we consider pure embeddings (by \cite{KjSh:455}). But it is not less natural to consider usual embeddings (remembering they, the (Abelian) groups under consideration, are reduced). The problem is that the invariant has been defined using divisibility, and so under non-pure embedding those seemed to be erased. Then in \cite{Sh:456} the non-existence of universals is proved restricting ourselves to $\lambda>2^{\aleph_0}$ and $(< \lambda)$-stable groups (see there). These restrictions hurt the generality of the theorem; because of the first requirement we lose some cardinals. The second requirement changes the class to one which is not established among Abelian group theorists (though to me it looks natural). Our aim was to eliminate those requirements, or show that they are necessary. Note that the present paper is mainly concerned essentially with results in ZFC, but they have roots in ``difficulties" in extending independence results thus providing a case for the \begin{Thesis} \label{0.2} Even if you do not like independence results you better look at them, as you will not even consider your desirable ZFC results when they are camouflaged by the litany of many independence results you can prove things. \end{Thesis} Of course, independence has interest {\em per se}; still for a given problem in general a solution in ZFC is for me preferable on an independence result. But if it gives a method of forcing (so relevant to a series of problems) the independence result is preferable (of course, I assume there are no other major differences; the depth of the proof would be of first importance to me). As occurs often in my papers lately, quotations of {\bf pcf} theory appear. This paper is also a case of \begin{Thesis} \label{0.3} Assumption of cases of not GCH at singular (more generally $pp\lambda> \lambda^+$) are ``good", ``helpful" assumptions; i.e. traditionally uses of GCH proliferate mainly not from conviction but as you can prove many theorems assuming $2^{\aleph_0}=\aleph_1$ but very few from $2^{\aleph_0}>\aleph_1$, but assuming $2^{\beth_\omega}>\beth^+_\omega$ is helpful in proving. \end{Thesis} Unfortunately, most results are only almost in ZFC as they use extremely weak assumptions from {\bf pcf}, assumptions whose independence is not know. So practically it is not tempting to try to remove them as they may be true, and it is unreasonable to try to prove independence results before independence results on {\bf pcf} will advance. In \S1 we give an explanation of the earlier difficulties: the problem (of the existence of universals for ${\frak K}^{rs(p)}$) is not like looking for ${\frak K}^{tr}$ (trees with $\omega+1$ levels) but for ${\frak K}^{tr}_{\langle\lambda_\alpha:\alpha< \omega\rangle}$ where \begin{description} \item[($\oplus$)] $\lambda^{\aleph_0}_n<\lambda_{n+1}<\mu$, $\lambda_n$ are regular and $\mu^+<\lambda=\lambda_\omega={\rm cf}(\lambda)<\mu^{\aleph_0}$ and ${\frak K}^{tr}_{\langle\lambda_n:n<\omega\rangle}$ is \[\{T:T\mbox{ a tree with $\omega+1$ levels, in level $n< \omega$ there are $\lambda_n$ elements}\}.\] \end{description} We also consider ${\frak K}^{tr}_{\langle\lambda_\alpha:\alpha\leq\omega \rangle}$, which is defined similarly but the level $\omega$ of $T$ is required to have $\lambda_\omega$ elements. \noindent For ${\frak K}^{rs(p)}$ this is proved fully, for ${\frak K}^{rtf}$ this is proved for the natural examples. \medskip In \S2 we define two such basic examples: one is ${\frak K}^{tr}_{\langle \lambda_\alpha:\alpha \le \omega \rangle}$, and the second is ${\frak K}^{fc}_{\langle \lambda_\alpha:\alpha\leq\omega\rangle}$. The first is a tree with $\omega+1$ levels; in the second we have slightly less restrictions. We have $\omega$ kinds of elements and a function from the $\omega$-th-kind to the $n$th kind. We can interpret a tree $T$ as a member of the second example: $P^T_\alpha = \{x:x \mbox{ is of level }\alpha\}$ and \[F_n(x) = y \quad\Leftrightarrow \quad x \in P^T_\omega\ \&\ y \in P^T_n\ \&\ y <_T x.\] For the second we recapture the non-existence theorems. But this is not one of the classes we considered originally. In \S3 we return to ${\frak K}^{rtf}$ (reduced torsion free Abelian groups) and prove the non-existence of universal ones in $\lambda$ if $2^{\aleph_0} < \mu^+<\lambda={\rm cf}(\lambda)<\mu^{\aleph_0}$ and an additional very weak set theoretic assumption (the consistency of its failure is not known). \medskip \noindent Note that (it will be proved in \cite{Sh:622}): \begin{description} \item[($\otimes$)] if $\lambda<2^{\aleph_0}$ then ${\frak K}^{rtf}_\lambda$ has no universal members. \end{description} \noindent Note: if $\lambda=\lambda^{\aleph_0}$ then ${\frak K}^{tr}_\lambda$ has universal member also ${\frak K}^{rs(p)}_\lambda$ (see \cite{Fu}) but not ${\frak K}^{rtf}_\lambda$ (see \cite[Ch IV, VI]{Sh:e}). \noindent We have noted above that for ${\frak K}^{rtf}_\lambda$ requiring $\lambda\geq 2^{\aleph_0}$ is reasonable: we can prove (i.e. in ZFC) that there is no universal member. What about ${\frak K}^{rs(p)}_\lambda$? By \S1 we should look at ${\frak K}^{tr}_{\langle\lambda_i:i\le\omega\rangle}$, $\lambda_\omega=\lambda<2^{\aleph_0}$, $\lambda_n<\aleph_0$. In \S4 we prove the consistency of the existence of universals for ${\frak K}^{tr}_{\langle\lambda_i:i \le\omega\rangle}$ when $\lambda_n\leq \omega$, $\lambda_\omega=\lambda< 2^{\aleph_0}$ but of cardinality $\lambda^+$; this is not the original problem but it seems to be a reasonable variant, and more seriously, it shoots down the hope to use the present methods of proving non-existence of universals. Anyhow this is ${\frak K}^{tr}_{\langle \lambda_i:i \le\omega\rangle}$ not ${\frak K}^{rs(p)}_{\lambda_\omega}$, so we proceed to reduce this problem to the previous one under a mild variant of MA. The intentions are to deal with ``there is universal of cardinality $\lambda$" in D\v{z}amonja Shelah \cite{DjSh:614}. The reader should remember that the consistency of e.g. \begin{quotation} {\em \noindent $2^{\aleph_0}>\lambda>\aleph_0$ and there is no $M$ such that $M\in {\frak K}^{rs(p)}$ is of cardinality $<2^{\aleph_0}$ and universal for ${\frak K}^{rs(p)}_\lambda$ } \end{quotation} is much easier to obtain, even in a wider context (just add many Cohen reals). \medskip As in \S4 the problem for ${\frak K}^{rs(p)}_\lambda$ was reasonably resolved for $\lambda<2^{\aleph_0}$ (and for $\lambda = \lambda^{\aleph_0}$, see \cite{KjSh:455}), we now, in \S5 turn to $\lambda>2^{\aleph_0}$ (and $\mu,\lambda_n$) as in $(\oplus)$ above. As in an earlier proof we use $\langle C_\delta: \delta\in S\rangle$ guessing clubs for $\lambda$ (see references or later here), so $C_\delta$ is a subset of $\delta$ (so the invariant depends on the representation of $G$ but this disappears when we divide by suitable ideal on $\lambda$). What we do is: rather than trying to code a subset of $C_\delta$ (for $\bar G=\langle G_i:i<\lambda\rangle$ a representation or filtration of the structure $G$ as the union of an increasing continuous sequence of structures of smaller cardinality) by an element of $G$, we do it, say, by some set $\bar x=\langle x_t:t\in {\rm Dom}(I)\rangle$, $I$ an ideal on ${\rm Dom}(I)$ (really by $\bar x/I$). At first glance if ${\rm Dom}(I)$ is infinite we cannot list {\em a priori} all possible such sequences for a candidate $H$ for being a universal member, as their number is $\ge\lambda^{\aleph_0}=\mu^{\aleph_0}$. But we can find a family \[{\cal F}\subseteq\{\langle x_t:t\in A\rangle:\ A\subseteq{\rm Dom}(I),\ A\notin I,\ x_t\in\lambda\}\] of cardinality $<\mu^{\aleph_0}$ such that for any $\bar{x}=\langle x_t:t\in {\rm Dom}(I)\rangle$, for some $\bar y\in {\cal F}$ we have $\bar y=\bar x\restriction {\rm Dom}(\bar y)$. \medskip As in \S3 there is such ${\cal F}$ except when some set theoretic statement related to {\bf pcf} holds. This statement is extremely strong, also in the sense that we do not know to prove its consistency at present. But again, it seems unreasonable to try to prove its consistency before the {\bf pcf} problem was dealt with. Of course, we may try to improve the combinatorics to avoid the use of this statement, but are naturally discouraged by the possibility that the {\bf pcf} statement can be proved in ZFC; thus we would retroactively get the non-existence of universals in ZFC. \medskip In \S6, under weak {\bf pcf} assumptions, we prove: if there is a universal member in ${\frak K}^{fc}_\lambda$ then there is one in ${\frak K}^{rs(p)}_\lambda$; so making the connection between the combinatorial structures and the algebraic ones closer. \medskip In \S7 we give other weak {\bf pcf} assumptions which suffice to prove non-existence of universals in ${\frak K}^x_{\langle\lambda_\alpha:\alpha \le\omega\rangle}$ (with $x$ one of the ``legal'' values): $\max{\rm pcf}\{\lambda_n:n<\omega\}=\lambda$ and ${\cal P}( \{\lambda_n:n<\omega\})/J_{<\lambda}\{\lambda_n:n<\omega\}$ is an infinite Boolean Algebra (and $(\oplus)$ holds, of course). \medskip In \cite{KjSh:409}, for singular $\lambda$ results on non-existence of universals (there on orders) can be gotten from these weak {\bf pcf} assumptions. \medskip In \S8 we get parallel results from, in general, more complicated assumptions. \medskip In \S9 we turn to a closely related class: the class of metric spaces with (one to one) continuous embeddings, similar results hold for it. We also phrase a natural criterion for deducing the non-existence of universals from one class to another. \medskip In \S10 we deal with modules and in \S11 we discuss the open problems of various degrees of seriousness. The sections are written in the order the research was done. \begin{notation} \label{0.4} Note that we deal with trees with $\omega+1$ levels rather than, say, with $\kappa+1$, and related situations, as those cases are quite popular. But inherently the proofs of \S1-\S3, \S5-\S9 work for $\kappa+1$ as well (in fact, {\bf pcf} theory is stronger). \noindent For a structure $M$, $\|M\|$ is its cardinality. \noindent For a model, i.e. a structure, $M$ of cardinality $\lambda$, where $\lambda$ is regular uncountable, we say that $\bar M$ is a representation (or filtration) of $M$ if $\bar M=\langle M_i:i<\lambda\rangle$ is an increasing continuous sequence of submodels of cardinality $<\lambda$ with union $M$. \noindent For a set $A$, we let $[A]^\kappa = \{B:B \subseteq A \mbox{ and } |B|=\kappa\}$. \noindent For a set $C$ of ordinals, $${\rm acc}(C)=\{\alpha\in C: \alpha=\sup(\alpha \cap C)\}, \mbox{(set of accumulation points)}$$ $$ {\rm nacc}(C)=C\setminus {\rm acc}(C) \ (=\mbox{ the set of non-accumulation points}). $$ We usually use $\eta$, $\nu$, $\rho$ for sequences of ordinals; let $\eta\vartriangleleft\nu$ means $\eta$ is an initial segment of $\nu$. Let ${\rm cov}(\lambda, \mu, \theta, \sigma)= \min\{|{\cal P}|: {\cal P}\subseteq [\lambda]^{<\mu}$, and for every $A\in [\lambda]^{<\theta}$ for some $\alpha< \sigma$ and $B_i\in {\cal P}$ for $i< \alpha$ we have $A\subseteq \bigcup\limits_{i< \alpha} B_i\}$. \noindent Remember that for an ordinal $\alpha$, e.g. a natural number, $\alpha=\{\beta:\beta<\alpha\}$. \end{notation} \begin{notation} \noindent ${\frak K}^{rs(p)}$ is the class of (Abelian) groups which are $p$-groups (i.e. $(\forall x\in G)(\exists n)[p^nx = 0]$) reduced (i.e. have no divisible non-zero subgroups) and separable (i.e. every cyclic pure subgroup is a direct summand). See \cite{Fu}. \noindent For $G\in{\frak K}^{rs(p)}$ define a norm $\|x\|=\inf\{2^{-n}: p^n \mbox{ divides } x\}$. Now every $G\in {\frak K}^{rs(p)}$ has a basic subgroup $B=\bigoplus\limits_{\scriptstyle n<\omega\atop\scriptstyle i<\lambda_n} {\Bbb Z} x^n_i$, where $x^n_i$ has order $p^{n+1}$, and every $x\in G$ can be represented as $\sum\limits_{\scriptstyle n<\omega\atop \scriptstyle i<\lambda_n} a^n_ix^n_i$, where for each $n$, $w_n(x)=\{i< \lambda_n:a^n_ix^n_i\ne 0\}$ is finite. \noindent ${\frak K}^{rtf}$ is the class of Abelian groups which are reduced and torsion free (i.e. $G \models nx = 0$, $n>0$\qquad $\Rightarrow\qquad x = 0$). \noindent For a group $G$ and $A\subseteq G$ let $\langle A\rangle_G$ be the subgroup of $G$ generated by $A$, we may omit the subscript $G$ if clear from the context. \noindent Group will mean an Abelian group, even if not stated explicitly. \noindent Let $H\subseteq_{pr} G$ means $H$ is a pure subgroup of $G$. \noindent Let $nG=\{nx: x\in G\}$ and let $G[n]=\{x\in G: nx=0\}$. \end{notation} \begin{notation} ${\frak K}$ will denote a class of structures with the same vocabulary, with a notion of embeddability, equivalently a notion $\leq_{{\frak K}}$ of submodel. \end{notation} \section{Their prototype is ${\frak K}^{tr}_{\langle \lambda_n:n<\omega \rangle}$ not ${\frak K}^{tr}$!} If we look for universal member in ${\frak K}^{rs(p)}_\lambda$, thesis \ref{0.1} suggests to us to think it is basically ${\frak K}^{tr}_\lambda$ (trees with $\omega+1$ levels, i.e. ${\frak K}^{tr}_{\lambda}$ is our prototype), a way followed in \cite{KjSh:455}, \cite{Sh:456}. But, as explained in the introduction, this does not give answer for the case of usual embedding for the family of all such groups. Here we show that for this case the thesis should be corrected. More concretely, the choice of the prototype means the choice of what we expect is the division of the possible classes. That is for a family of classes a choice of a prototype assert that we believe that they all behave in the same way. We show that looking for a universal member $G$ in ${\frak K}^{rs(p)}_\lambda$ is like looking for it among the $G$'s with density $\le\mu$ ($\lambda,\mu$, as usual, as in $(\oplus)$ from \S0). For ${\frak K}^{rtf}_\lambda$ we get weaker results which still cover the examples usually constructed, so showing that the restrictions in \cite{KjSh:455} (to pure embeddings) and \cite{Sh:456} (to $(<\lambda)$-stable groups) were natural. \begin{Proposition} \label{1.1} Assume that $\mu=\sum\limits_{n<\omega}\lambda_n=\lim\sup\limits_n\lambda_n$, $\mu\le\lambda\le\mu^{\aleph_0}$, and $G$ is a reduced separable $p$-group such that \[|G|=\lambda\quad\mbox{ and }\quad\lambda_n(G)=:\dim((p^n G)[p]/ (p^{n+1}G)[p])\le\mu\] (this is a vector space over ${\Bbb Z}/p {\Bbb Z}$, hence the dimension is well defined). \\ {\em Then} there is a reduced separable $p$-group $H$ such that $|H|=\lambda$, $H$ extends $G$ and $(p^nH)[p]/(p^{n+1}H)[p]$ is a group of dimension $\lambda_n$ (so if $\lambda_n\geq \aleph_0$, this means cardinality $\lambda_n$). \end{Proposition} \begin{Remark} \label{1.1A} So for $H$ the invariants from \cite{KjSh:455} are trivial. \end{Remark} \proof (See Fuchs \cite{Fu}). We can find $z^n_i$ (for $n<\omega$, $i<\lambda_n(G)\le\mu$) such that: \begin{description} \item[(a)] $z^n_i$ has order $p^n$, \item[(b)] $B=\sum\limits_{n,i}\langle z^n_i \rangle_G$ is a direct sum, \item[(c)] $B$ is dense in $G$ in the topology induced by the norm \[\|x\|=\min\{2^{-n}:p^n \mbox{ divides } x \mbox{ in } G\}.\] \end{description} For each $n<\omega$ and $i<\lambda_n(G)$ ($\le\mu$) choose $\eta^n_i\in \prod\limits_{m<\omega}\lambda_m$, pairwise distinct such that for $(n^1,i^1) \neq (n^2,i^2)$ for some $n(*)$ we have: \[\lambda_n \ge \lambda_{n(*)}\qquad \Rightarrow\qquad \eta^{n^1}_{i^1}(n) \neq \eta^{n^2}_{i^2}(n).\] Let $H$ be generated by $G$, $x^m_i$ ($i<\lambda_m$, $m<\omega$), $y^{n,k}_i$ ($i<\lambda_n$, $n<\omega$, $n\le k<\omega)$ freely except for: \begin{description} \item[($\alpha$)] the equations of $G$, \item[($\beta$)] $y^{n,n}_i = z^n_i$, \item[($\gamma$)] $py^{n,k+1}_i - y^{n,k}_i = x^k_{\eta^n_i(k)}$, \item[($\delta$)] $p^{n+1}x^n_i = 0$, \item[($\varepsilon$)] $p^{k+1}y^{n,k}_i = 0$. \end{description} Now check. \hfill$\square_{\ref{1.1}}$ \begin{Definition} \label{1.2} \begin{enumerate} \item ${\bf t}$ denotes a sequence $\langle t_i:i<\omega\rangle$, $t_i$ a natural number $>1$. \item For a group $G$ we define \[G^{[{\bf t}]}=\{x\in G:\bigwedge_{j<\omega}[x\in (\prod_{i<j} t_i) G]\}.\] \item We can define a semi-norm $\|-\|_{{\bf t}}$ on $G$ \[\|x\|_{{\bf t}}=\min\{2^{-i}:x\in (\prod_{j<i} t_j)G\}\] and so the semi-metric \[d_{{\bf t}}(x,y)=\|x-y\|_{{\bf t}}.\] \end{enumerate} \end{Definition} \begin{Remark} \label{1.2A} So, if $\|-\|_{{\bf t}}$ is a norm, $G$ has a completion under $\|-\|_{{\bf t}}$, which we call $\|-\|_{{\bf t}}$-completion; if ${\bf t}=\langle i!:i<\omega \rangle$ we refer to $\|-\|_{{\bf t}}$ as $\Bbb Z$-adic norm, and this induces $\Bbb Z$-adic topology, so we can speak of $\Bbb Z$-adic completion. \end{Remark} \begin{Proposition} \label{1.3} Suppose that \begin{description} \item[($\otimes_0$)] $\mu=\sum\limits_n\lambda_n$ and $\mu\le\lambda\le \mu^{\aleph_0}$ for simplicity, $2<2\cdot\lambda_n\le\lambda_{n+1}$ (maybe $\lambda_n$ is finite!), \item[($\otimes_1$)] $G$ is a torsion free group, $|G|=\lambda$, and $G^{[{\bf t}]}=\{0\}$, \item[($\otimes_2$)] $G_0\subseteq G$, $G_0$ is free and $G_0$ is ${\bf t}$-dense in $G$ (i.e. in the topology induced by the metric $d_{{\bf t}}$), where ${\bf t}$ is a sequence of primes. \end{description} {\em Then} there is a torsion free group $H$, $G\subseteq H$, $H^{[{\bf t}]} =\{0\}$, $|H|=\lambda$ and, under $d_{{\bf t}}$, $H$ has density $\mu$. \end{Proposition} \proof Let $\{x_i:i<\lambda\}$ be a basis of $G_0$. Let $\eta_i\in \prod\limits_{n<\omega} \lambda_n$ for $i<\mu$ be distinct such that $\eta_i(n+1)\geq \lambda_n$ and \[i\ne j\qquad \Rightarrow\qquad (\exists m)(\forall n)[m \le n \quad \Rightarrow\quad \eta_i(n) \ne \eta_j(n)].\] Let $H$ be generated by \[G,\ \ x^m_i \mbox{ (for $i<\lambda_m$, $m<\omega$), }\ y^n_i \mbox{ (for $i<\mu$, $n<\omega$)}\] freely except for \begin{description} \item[(a)] the equations of $G$, \item[(b)] $y^0_i = x_i$, \item[(c)] $t_n\, y^{n+1}_i + y^n_i = x^n_{\eta_i(n)}$. \end{description} \medskip \noindent{\bf Fact A}\hspace{0.15in} $H$ extends $G$ and is torsion free. \noindent [Why? As $H$ can be embedded into the divisible hull of $G$.] \medskip \noindent{\bf Fact B}\hspace{0.15in} $H^{[{\bf t}]}= \{0\}$. \proof Let $K$ be a countable pure subgroup of $H$ such that $K^{[{\bf t}]}\ne \{0\}$. Now without loss of generality $K$ is generated by \begin{description} \item[(i)] $K_1\subseteq G\cap\mbox{ [the $d_{{\bf t}}$--closure of $\langle x_i: i\in I\rangle_G]$]}$, where $I$ is a countable infinite subset of $\lambda$ and $K_1\supseteq\langle x_i:i\in I\rangle_G$, \item[(ii)] $y^m_i$, $x^n_j$ for $i\in I$, $m<\omega$ and $(n,j)\in J$, where $J\subseteq \omega\times \lambda$ is countable and \[i\in I,\ n<\omega\qquad\Rightarrow\qquad (n,\eta_i(n))\in J.\] \end{description} Moreover, the equations holding among those elements are deducible from the equations of the form \begin{description} \item[(a)$^-$] equations of $K_1$, \item[(b)$^-$] $y^0_i=x_i$ for $i \in I$, \item[(c)$^-$] $t_n\,y^{n+1}_i+y^n_i=x^n_{\eta_i(n)}$ for $i\in I,n<\omega$. \end{description} \noindent We can find $\langle k_i:i<\omega\rangle$ such that \[[n\ge k_i\ \&\ n\ge k_j\ \&\ i \ne j\qquad \Rightarrow\qquad \eta_i(n)\ne \eta_j(n)].\] Let $y \in K\setminus\{0\}$. Then for some $j$, $y\notin (\prod\limits_{i<j} t_i)G$, so for some finite $I_0\subseteq I$ and finite $J_0\subseteq J$ and \[y^* \in\langle\{x_i:i\in I_0\}\cup\{x^n_\alpha:(n,\alpha)\in J_0\} \rangle_K\] we have $y-y^*\in(\prod\limits_{i<j} t_i) G$. Without loss of generality $J_0 \cap\{(n,\eta_i(n)):i\in I,\ n\ge k_i\}=\emptyset$. Now there is a homomorphism $\varphi$ from $K$ into the divisible hull $K^{**}$ of \[K^* = \langle\{x_i:i\in I_0\}\cup\{x^n_j:(n,j)\in J_0\}\rangle_G\] such that ${\rm Rang}(\varphi)/K^*$ is finite. This is enough. \medskip \noindent{\bf Fact C}\hspace{0.15in} $H_0=:\langle x^n_i:n<\omega,i<\lambda_n \rangle_H$ is dense in $H$ by $d_{{\bf t}}$. \proof Straight as each $x_i$ is in the $d_{{\bf t}}$-closure of $H_0$ inside $H$. \medskip Noting then that we can increase the dimension easily, we are done. \hfill$\square_{\ref{1.3}}$ \section{On structures like $(\prod\limits_n \lambda_n,E_m)_{m<\omega}$, $\eta E_m \nu =: \eta(m)=\nu(m)$} \begin{Discussion} \label{2.1} We discuss the existence of universal members in cardinality $\lambda$, $\mu^+<\lambda<\mu^{\aleph_0}$, for certain classes of groups. The claims in \S1 indicate that the problem is similar not to the problem of the existence of a universal member in ${\frak K}^{tr}_\lambda$ (the class of trees with $\lambda$ nodes, $\omega+1$ levels) but to the one where the first $\omega$ levels, are each with $<\mu$ elements. We look more carefully and see that some variants are quite different. The major concepts and Lemma (\ref{2.3}) are similar to those of \S3, but easier. Since detailed proofs are given in \S3, here we give somewhat shorter proofs. \end{Discussion} \begin{Definition} \label{2.2} For a sequence $\bar\lambda=\langle\lambda_i:i\le\delta\rangle$ of cardinals we define: \begin{description} \item[(A)] ${\frak K}^{tr}_{\bar \lambda}=\{T:\,T$ is a tree with $\delta +1$ levels (i.e. a partial order such that \qquad\quad for $x\in T$, ${\rm lev}_T(x)=:{\rm otp}(\{y:y<x\})$ is an ordinal $\le\delta$) such \qquad\quad that:\quad ${\rm lev}_i(T)=:\{x\in T:{\rm lev}_T(x)=i\}$ has cardinality $\le\lambda_i\}$, \item[(B)] ${\frak K}^{fc}_{\bar\lambda}=\{M:\,M=(|M|,P_i,F_i)_{i\le\delta}$, $|M|$ is the disjoint union of \qquad\quad $\langle P_i:i\le\delta\rangle$, $F_i$ is a function from $P_\delta$ to $P_i$, $\|P_i\|\le\lambda_i$, \qquad\quad $F_\delta$ is the identity (so can be omitted)$\}$, \item[(C)] If $[i\le\delta\quad \Rightarrow\quad \lambda_i=\lambda]$ then we write $\lambda$, $\delta+1$ instead of $\langle\lambda_i:i\le\delta\rangle$. \end{description} \end{Definition} \begin{Definition} \label{2.2A} Embeddings for ${\frak K}^{tr}_{\bar\lambda}$, ${\frak K}^{fc}_{\bar\lambda}$ are defined naturally: for ${\frak K}^{tr}_{\bar\lambda}$ embeddings preserve $x<y$, $\neg x<y$, ${\rm lev}_T(x)=\alpha$; for ${\frak K}^{fc}_{\bar\lambda}$ embeddings are defined just as for models. If $\delta^1=\delta^2=\delta$ and $[i<\delta\quad\Rightarrow\quad\lambda^1_i \le\lambda^2_i]$ and $M^\ell\in{\frak K}^{fc}_{\bar\lambda^\ell}$, (or $T^\ell \in{\frak K}^{tr}_{\bar\lambda^\ell}$) for $\ell=1,2$, then an embedding of $M^1$ into $M^2$ ($T^1$ into $T^2$) is defined naturally. \end{Definition} \begin{Lemma} \label{2.3} Assume $\bar\lambda=\langle\lambda_i:i\le\delta\rangle$ and $\theta$, $\chi$ satisfy (for some $\bar C$): \begin{description} \item[(a)] $\lambda_\delta$, $\theta$ are regular, $\bar C=\langle C_\alpha: \alpha\in S\rangle$, $S\subseteq\lambda=:\lambda_\delta$, $C_\alpha\subseteq \alpha$, for every club $E$ of $\lambda$ for some $\alpha$ we have $C_\alpha \subseteq E$, $\lambda_\delta<\chi\le |C_\alpha|^\theta$ and ${\rm otp}(C_\alpha) \ge\theta$, \item[(b)] $\lambda_i\le\lambda_\delta$, \item[(c)] there are $\theta$ pairwise disjoint sets $A\subseteq\delta$ such that $\prod\limits_{i\in A}\lambda_i\ge\lambda_\delta$. \end{description} {\em Then} \begin{description} \item[($\alpha$)] there is no universal member in ${\frak K}^{fc}_{\bar \lambda}$;\quad moreover \item[($\beta$)] if $M_\alpha\in {\frak K}^{fc}_{\bar\lambda}$ or even $M_\alpha\in {\frak K}^{fc}_{\lambda_\delta}$ for $\alpha<\alpha^*<\chi$ {\em then} some $M\in {\frak K}^{fc}_{\bar\lambda}$ cannot be embedded into any $M_\alpha$. \end{description} \end{Lemma} \begin{Remark} \label{2.3A} Note that clause $(\beta)$ is relevant to our discussion in \S1: the non-universality is preserved even if we increase the density and, also, it is witnessed even by non-embeddability in many models. \end{Remark} \proof Let $\langle A_\varepsilon:\varepsilon<\theta\rangle$ be as in clause (c) and let $\eta^\varepsilon_\alpha\in\prod\limits_{i\in A_\varepsilon} \lambda_i$ for $\alpha<\lambda_\delta$ be pairwise distinct. We fix $M_\alpha \in {\frak K}^{fc}_{\lambda_\delta}$ for $\alpha<\alpha^*<\chi$. \noindent For $M\in {\frak K}^{fc}_{\bar\lambda}$, let $\bar M=(|M|,P^M_i, F^M_i)_{i\le\delta}$ and let $\langle M_\alpha: \alpha< \lambda_\delta\rangle$ be a representation (=filtration) of $M$; for $\alpha\in S$, $x\in P^M_\delta$, let \[\begin{ALIGN} {\rm inv}(x,C_\alpha;\bar M)=\big\{\beta\in C_\alpha:&\mbox{for some }\varepsilon <\theta\mbox{ and } y\in M_{\min(C_\alpha\setminus (\beta+1))}\\ &\mbox{we have }\ \bigwedge\limits_{i\in A_\varepsilon} F^M_i(x)=F^M_i(y)\\ &\mbox{\underbar{but} there is no such } y\in M_\beta\big\}. \end{ALIGN}\] \[{\rm Inv}(C_\alpha,\bar M)=\{{\rm inv}(x,C_\alpha,\bar M):x\in P^M_\delta\}.\] \[{\rm INv}(\bar M,\bar C)=\langle{\rm Inv}(C_\alpha,\bar M):\alpha\in S\rangle.\] \[{\rm INV}(\bar M,\bar C)={\rm INv}(\bar M,\bar C)/{\rm id}^a(\bar C).\] Recall that \[{\rm id}^a(\bar C)=\{T\subseteq\lambda:\mbox{ for some club $E$ of $\lambda$ for no $\alpha\in T$ is $C_\alpha\subseteq E$}\}.\] The rest should be clear (for more details see proofs in \S3), noticing \begin{Fact} \label{2.3B} \begin{enumerate} \item ${\rm INV}(\bar M,\bar C)$ is well defined, i.e. if $\bar M^1$, $\bar M^2$ are representations (=filtrations) of $M$ then ${\rm INV}(\bar M^1,\bar C)={\rm INV}(\bar M^2,\bar C)$. \item ${\rm Inv}(C_\alpha,\bar M)$ has cardinality $\le\lambda$. \item ${\rm inv}(x,C_\alpha;\bar M)$ is a subset of $C_\alpha$ of cardinality $\le \theta$. \end{enumerate} \end{Fact} \hfill$\square_{\ref{2.3}}$ \begin{Conclusion} \label{2.4} If $\mu=\sum\limits_{n<\omega}\lambda_n$ and $\lambda^{\aleph_0}_n< \lambda_{n+1}$ and $\mu^+<\lambda_\omega={\rm cf}(\lambda_\omega)<\mu^{\aleph_0}$, {\em then} in ${\frak K}^{fc}_{\langle\lambda_\alpha:\alpha\le\omega\rangle}$ there is no universal member and even in ${\frak K}^{fc}_{\langle \lambda_\omega:\alpha\le\omega\rangle}$ we cannot find a member universal for it. \end{Conclusion} \proof Should be clear or see the proof in \S3. \hfill$\square_{\ref{2.4}}$ \section{Reduced torsion free groups: Non-existence of universals} We try to choose torsion free reduced groups and define invariants so that in an extension to another such group $H$ something survives. To this end it is natural to stretch ``reduced" near to its limit. \begin{Definition} \label{3.1} \begin{enumerate} \item ${\frak K}^{tf}$ is the class of torsion free (abelian) groups. \item ${\frak K}^{rtf}=\{G\in {\frak K}^{tf}:{\Bbb Q}$ is not embeddable into $G$ (i.e. $G$ is reduced)$\}$. \item ${\bf P}^*$ denotes the set of primes. \item For $x\in G$, ${\bf P}(x,G)=:\{p\in{\bf P}^*: \bigwedge\limits_n x\in p^n G\}$. \item ${\frak K}^x_\lambda=\{G\in{\frak K}^x:\|G\|=\lambda\}$. \item If $H\in {\frak K}^{rtf}_\lambda$, we say $\bar H$ is a representation or filtration of $H$ if $\bar H=\langle H_\alpha:\alpha<\lambda\rangle$ is increasing continuous and $H=\bigcup\limits_{\alpha<\lambda} H_\alpha$, $H\in {\frak K}^{rtf}$ and each $H_\alpha$ has cardinality $<\lambda$. \end{enumerate} \end{Definition} \begin{Proposition} \label{3.2} \begin{enumerate} \item If $G\in {\frak K}^{rtf}$, $x\in G\setminus\{0\}$, $Q\cup{\bf P}(x,G) \subsetneqq{\bf P}^*$, $G^+$ is the group generated by $G,y,y_{p,\ell}$ ($\ell <\omega$, $p\in Q$) freely, except for the equations of $G$ and \[y_{p,0}=y,\quad py_{p,\ell+1}=y_{p,\ell}\quad \mbox{ and }\quad y_{p,\ell}=z\mbox{ when } z\in G,p^\ell z=x\] {\em then} $G^+\in {\frak K}^{rtf}$, $G\subseteq_{pr}G^+$ (pure extension). \item If $G_i\in {\frak K}^{rtf}$ ($i<\alpha$) is $\subseteq_{pr}$-increasing {\em then} $G_i\subseteq_{pr}\bigcup\limits_{j<\alpha}G_j\in{\frak K}^{rtf}$ for every $i<\alpha$. \end{enumerate} \end{Proposition} The proof of the following lemma introduces a method quite central to this paper. \begin{Lemma} \label{3.3} Assume that \begin{description} \item[$(*)^1_\lambda$] $2^{\aleph_0}+\mu^+<\lambda={\rm cf}(\lambda)< \mu^{\aleph_0}$, \item[$(*)^2_\lambda$] for every $\chi<\lambda$, there is $S\subseteq [\chi]^{\le\aleph_0}$, such that: \begin{description} \item[(i)] $|S|<\lambda$, \item[(ii)] if $D$ is a non-principal ultrafilter on $\omega$ and $f:D \longrightarrow\chi$ {\em then} for some $a\in S$ we have \[\bigcap \{X\in D:f(X)\in a\}\notin D.\] \end{description} \end{description} {\em Then} \begin{description} \item[($\alpha$)] in ${\frak K}^{rtf}_\lambda$ there is no universal member (under usual embeddings (i.e. not necessarily pure)), \item[($\beta$)] moreover, \underbar{for any} $G_i\in {\frak K}^{rtf}_\lambda$, for $i<i^*<\mu^{\aleph_0}$ \underbar{there is} $G\in {\frak K}^{rtf}_\lambda$ not embeddable into any one of $G_i$. \end{description} \end{Lemma} Before we prove \ref{3.3} we consider the assumptions of \ref{3.3} in \ref{3.4}, \ref{3.5}. \begin{Claim} \label{3.4} \begin{enumerate} \item In \ref{3.3} we can replace $(*)^1_\lambda$ by \begin{description} \item[$(**)^1_\lambda$ (i)] $2^{\aleph_0}<\mu<\lambda={\rm cf}(\lambda)< \mu^{\aleph_0}$, \item[\qquad(ii)] there is $\bar C=\langle C_\delta:\delta\in S^*\rangle$ such that $S^*$ is a stationary subset of $\lambda$, each $C_\delta$ is a subset of $\delta$ with ${\rm otp}(C_\delta)$ divisible by $\mu$, $C_\delta$ closed in $\sup(C_\delta)$ (which normally $\delta$, but not necessarily so) and \[(\forall\alpha)[\alpha\in {\rm nacc}(C_\delta)\quad \Rightarrow\quad {\rm cf}(\alpha) >2^{\aleph_0}]\] (where ${\rm nacc}$ stands for ``non-accumulation points''), and such that $\bar C$ guesses clubs of $\lambda$ (i.e. for every club $E$ of $\lambda$, for some $\delta\in S^*$ we have $C_\delta\subseteq E$) and $[\delta\in S^*\quad \Rightarrow\quad {\rm cf}(\delta)=\aleph_0]$. \end{description} \item In $(*)^1_\lambda$ and in $(*)^2_\lambda$, without loss of generality $(\forall\theta<\mu)[\theta^{\aleph_0}<\mu]$ and ${\rm cf}(\mu)=\aleph_0$. \end{enumerate} \end{Claim} \proof \ \ \ 1) This is what we actually use in the proof (see below). \noindent 2) Replace $\mu$ by $\mu'=\min\{\mu_1:\mu^{\aleph_0}_1\ge\mu$ (equivalently $\mu^{\aleph_0}_1=\mu^{\aleph_0}$)$\}$. \hfill$\square_{\ref{3.4}}$ Compare to, say, \cite{KjSh:447}, \cite{KjSh:455}; the new assumption is $(*)^2_\lambda$, note that it is a very weak assumption, in fact it might be that it is always true. \begin{Claim} \label{3.5} Assume that $2^{\aleph_0}<\mu<\lambda<\mu^{\aleph_0}$ and $(\forall \theta< \mu)[\theta^{\aleph_0}<\mu]$ (see \ref{3.4}(2)). Then each of the following is a sufficient condition to $(*)^2_\lambda$: \begin{description} \item[($\alpha$)] $\lambda<\mu^{+\omega_1}$, \item[($\beta$)] if ${\frak a}\subseteq{\rm Reg}\cap\lambda\setminus\mu$ and $|{\frak a}|\le 2^{\aleph_0}$ then we can find $h:{\frak a}\longrightarrow \omega$ such that: \[\lambda>\sup\{\max{\rm pcf}({\frak b}):{\frak b}\subseteq {\frak a}\mbox{ countable, and $h\restriction {\frak b}$ constant}\}.\] \end{description} \end{Claim} \proof Clause $(\alpha)$ implies Clause $(\beta)$: just use any one-to-one function $h:{\rm Reg}\cap\lambda\setminus\mu\longrightarrow\omega$. \smallskip Clause $(\beta)$ implies (by \cite[\S6]{Sh:410} + \cite[\S2]{Sh:430}) that for $\chi<\lambda$ there is $S\subseteq [\chi]^{\aleph_0}$, $|S|<\lambda$ such that for every $Y\subseteq\chi$, $|Y|=2^{\aleph_0}$, we can find $Y_n$ such that $Y=\bigcup\limits_{n<\omega} Y_n$ and $[Y_n]^{\aleph_0} \subseteq S$. (Remember: $\mu>2^{\aleph_0}$.) Without loss of generality (as $2^{\aleph_0} < \mu < \lambda$): \begin{description} \item[$(*)$] $S$ is downward closed. \end{description} So if $D$ is a non-principal ultrafilter on $\omega$ and $f:D\longrightarrow \chi$ then letting $Y={\rm Rang}(f)$ we can find $\langle Y_n:n<\omega\rangle$ as above. Let $h:D\longrightarrow\omega$ be defined by $h(A)=\min\{n:f(A)\in Y_n\}$. So \[X\subseteq D\ \ \&\ \ |X|\le\aleph_0\ \ \&\ \ h\restriction X \mbox{ constant }\Rightarrow\ f''(X)\in S\quad\mbox{(remember $(*)$)}.\] Now for each $n$, for some countable $X_n\subseteq D$ (possibly finite or even empty) we have: \[h \restriction X_n\ \mbox{ is constantly } n,\] \[\ell <\omega \ \&\ (\exists A\in D)(h(A)=n\ \&\ \ell\notin A) \Rightarrow (\exists B\in X_n)(\ell\notin B).\] Let $A_n=:\bigcap\{A:A\in X_n\}=\bigcap\{A:A\in D\mbox{ and } h(X)=n\}$. If the desired conclusion fails, then $\bigwedge\limits_{n<\omega}A_n\in D$. So \[(\forall A)[A\in D\quad \Leftrightarrow\quad\bigvee_{n<\omega} A\supseteq A_n].\] So $D$ is generated by $\{A_n:n<\omega\}$ but then $D$ cannot be a non-principal ultrafilter. \hfill$\square_{\ref{3.5}}$ \begin{Remark} The case when $D$ is a principal ultrafilter is trivial. \end{Remark} \proof of Lemma \ref{3.3} Let $\bar C=\langle C_\delta:\delta\in S^*\rangle$ be as in $(**)^1_{\bar \lambda}$ (ii) from \ref{3.4} (for \ref{3.4}(1) its existence is obvious, for \ref{3.3} - use \cite[VI, old III 7.8]{Sh:e}). Let us suppose that $\bar A=\langle A_\delta:\delta\in S^*\rangle$, $A_\delta \subseteq{\rm nacc}(C_\delta)$ has order type $\omega$ ($A_\delta$ like this will be chosen later) and let $\eta_\delta$ enumerate $A_\delta$ increasingly. Let $G_0$ be freely generated by $\{x_i:i<\lambda\}$. Let $R$ be \[\begin{ALIGN} \big\{\bar a: &\bar a=\langle a_n:n<\omega\rangle\mbox{ is a sequence of pairwise disjoint subsets of } {\bf P}^*,\\ &\mbox{with union }{\bf P}^* \mbox{ for simplicity, such that}\\ &\mbox{for infinitely many }n,\ a_n\ne\emptyset\big\}. \end{ALIGN}\] Let $G$ be a group generated by \[G_0 \cup \{y^{\alpha,n}_{\bar a},z^{\alpha,n}_{\bar a,p}:\ \alpha<\lambda, \ \bar a\in R,\ n<\omega,\ p \mbox{ prime}\}\] freely except for: \begin{description} \item[(a)] the equations of $G_0$, \item[(b)] $pz^{\alpha,n+1}_{\bar a,p}=z^{\alpha,n}_{\bar a,p}$ when $p\in a_n$, $\alpha<\lambda$, \item[(c)] $z^{\delta,0}_{\bar a,p}=y^{\delta,n}_{\bar a}- x_{\eta_\delta(n)}$ when $p\in a_n$ and $\delta\in S^*$. \end{description} Now $G\in {\frak K}^{rtf}_\lambda$ by inspection. \medskip \noindent Before continuing the proof of \ref{3.3} we present a definition and some facts. \begin{Definition} \label{3.7} For a representation $\bar H$ of $H\in {\frak K}^{rtf}_\lambda$, and $x\in H$, $\delta\in S^*$ let \begin{enumerate} \item ${\rm inv}(x,C_\delta;\bar H)=:\{\alpha\in C_\delta:$ for some $Q\subseteq {\bf P}^*$, there is $y\in H_{\min[C_\delta\setminus(\alpha+1)]}$ such that $Q\subseteq{\bf P}(x-y,H)$ but for no $y\in H_\alpha$ is $Q\subseteq{\bf P}(x-y,H)\}$ (so ${\rm inv}(x,C_\delta;\bar H)$ is a subset of $C_\delta$ of cardinality $\le 2^{\aleph_0}$). \item ${\rm Inv}^0(C_\delta,\bar H)=:\{{\rm inv}(x,C_\delta;\bar H):x\in \bigcup\limits_i H_i\}$. \item ${\rm Inv}^1(C_\delta,\bar H)=:\{a:a\subseteq C_\delta$ countable and for some $x\in H$, $a\subseteq{\rm inv}(x,C_\delta;\bar H)\}$. \item ${\rm INv}^\ell(\bar H,\bar C)=:{\rm Inv}^\ell(H,\bar H,\bar C)=:\langle {\rm Inv}^\ell(C_\delta;\bar H):\delta\in S^*\rangle$ for $\ell\in\{0,1\}$. \item ${\rm INV}^\ell(H,\bar C)=:{\rm INv}^\ell(H,\bar H,\bar C)/{\rm id}^a(\bar C)$, where \[{\rm id}^a(\bar C)=:\{T\subseteq\lambda:\mbox{ for some club $E$ of $\lambda$ for no $\delta\in T$ is }C_\delta\subseteq E\}.\] \item If $\ell$ is omitted, $\ell = 0$ is understood. \end{enumerate} \end{Definition} \begin{Fact} \label{3.8} \begin{enumerate} \item ${\rm INV}^\ell(H,\bar C)$ is well defined. \item The $\delta$-th component of ${\rm INv}^\ell(\bar H,\bar C)$ is a family of $\le\lambda$ subsets of $C_\delta$ each of cardinality $\le 2^{\aleph_0}$ and if $\ell=1$ each member is countable and the family is closed under subsets. \item {\em If} $G_i\in{\frak K}^{rtf}_\lambda$ for $i<i^*$, $i^*< \mu^{\aleph_0}$, $\bar G^i=\langle\bar G_{i,\alpha}:\alpha<\lambda\rangle$ is a representation of $G_i$, {\em then} we can find $A_\delta\subseteq{\rm nacc}(C_\delta)$ of order type $\omega$ such that: $i<i^*$, $\delta\in S^*\qquad \Rightarrow$\qquad for no $a$ in the $\delta$-th component of ${\rm INv}^\ell(G_i,\bar G^i,\bar C)$ do we have $|a \cap A_\delta|\ge\aleph_0$. \end{enumerate} \end{Fact} \proof Straightforward. (For (3) note ${\rm otp}(C_\delta)\ge\mu$, so there are $\mu^{\aleph_0}>\lambda$ pairwise almost disjoint subsets of $C_\delta$ each of cardinality $\aleph_0$ and every $A\in{\rm Inv}(C_\delta,\bar G^i)$ disqualifies at most $2^{\aleph_0}$ of them.) \hfill$\square_{\ref{3.8}}$ \begin{Fact} \label{3.9} Let $G$ be as constructed above for $\langle A_\delta:\delta\in S^*\rangle,A_\delta\subseteq{\rm nacc}(C_\delta)$, ${\rm otp}(A_\delta)=\omega$ (where $\langle A_\delta:\delta\in S^*\rangle$ are chosen as in \ref{3.8}(3) for the sequence $\langle G_i:i<i^* \rangle$ given for proving \ref{3.3}, see $(\beta)$ there). \noindent Assume $G \subseteq H\in {\frak K}^{rtf}_\lambda$ and $\bar H$ is a filtration of $H$. {\em Then} \[\begin{array}{rr} B=:\big\{\delta:A_\delta\mbox{ has infinite intersection with some}&\ \\ a\in{\rm Inv}(C_\delta,\bar H)\big\}&=\ \lambda\ \mod\ {\rm id}^a(\bar C). \end{array}\] \end{Fact} \proof We assume otherwise and derive a contradiction. Let for $\alpha <\lambda$, $S_\alpha\subseteq [\alpha]^{\le \aleph_0}$, $|S_\alpha|<\lambda$ be as guaranteed by $(*)^2_\lambda$. Let $\chi>2^\lambda$, ${\frak A}_\alpha\prec (H(\chi),\in,<^*_\chi)$ for $\alpha<\lambda$ increasing continuous, $\|{\frak A}_\alpha\|<\lambda$, $\langle {\frak A}_\beta:\beta\le\alpha\rangle\in {\frak A}_{\alpha+1}$, ${\frak A}_\alpha\cap\lambda$ an ordinal and: \[\langle S_\alpha:\alpha<\lambda\rangle,\ G,\ H,\ \bar C,\ \langle A_\delta:\delta\in S^* \rangle,\ \bar H,\ \langle x_i, y^\delta_{\bar a}, z^{\delta,n}_{\bar a,p}:\;i,\delta,\bar a,n,p \rangle\] all belong to ${\frak A}_0$ and $2^{\aleph_0}+1\subseteq {\frak A}_0$. Then $E=\{\delta<\lambda:{\frak A}_\delta \cap\lambda=\delta\}$ is a club of $\lambda$. Choose $\delta\in S^* \cap E\setminus B$ such that $C_\delta \subseteq E$. (Why can we? As to ${\rm id}^a(\bar C)$ belong all non stationary subsets of $\lambda$, in particular $\lambda\setminus E$, and $\lambda\setminus S^*$ and $B$, but $\lambda\notin {\rm id}^a(\bar C)$.) Remember that $\eta_\delta$ enumerates $A_\delta$ (in the increasing order). For each $\alpha\in A_\delta$ (so $\alpha\in E$ hence ${\frak A}_\alpha \cap \lambda=\alpha$ but $\bar H\in {\frak A}_\alpha$ hence $H\cap {\frak A}_\alpha= H_\alpha$) and $Q\subseteq{\bf P}^*$ choose, if possible, $y_{\alpha,Q}\in H_\alpha$ such that: \[Q\subseteq{\bf P}(x_\alpha-y_{\alpha,Q},H).\] Let $I_\alpha=:\{Q\subseteq{\bf P}^*:y_{\alpha,Q}$ well defined$\}$. Note (see \ref{3.4} $(**)^1_\lambda$ and remember $\eta_\delta(n)\in A_\delta\subseteq {\rm nacc}(C_\delta)$) that ${\rm cf}(\alpha)>2^{\aleph_0}$ (by (ii) of \ref{3.4} $(**)^1_\lambda$) and hence for some $\beta_\alpha<\alpha$, \[\{ y_{\alpha,Q}:Q\in I_\alpha\}\subseteq H_{\beta_\alpha}.\] Now: \begin{description} \item[$\otimes_1$] $I_\alpha$ is downward closed family of subsets of ${\bf P}^*$, ${\bf P}^*\notin I_\alpha$ for $\alpha \in A_\delta$. \end{description} [Why? See the definition for the first phrase and note also that $H$ is reduced for the second phrase.] \begin{description} \item[$\otimes_2$] $I_\alpha$ is closed under unions of two members (hence is an ideal on ${\bf P}^*$). \end{description} [Why? If $Q_1,Q_2\in I_\alpha$ then (as $x_\alpha\in G\subseteq H$ witnesses this): \[\begin{ALIGN} ({\cal H}(\chi),\in,<^*_\chi)\models &(\exists x)(x\in H\ \&\ Q_1\subseteq{\bf P}(x- y_{\alpha,Q_1},H)\ \&\\ &Q_2\subseteq{\bf P}(x-y_{\alpha,Q_2},H)). \end{ALIGN}\] All the parameters are in ${\frak A}_\alpha$ so there is $y\in {\frak A}_\alpha\cap H$ such that \[Q_1\subseteq{\bf P}(y-y_{\alpha,Q_1},H)\quad\mbox{ and }\quad Q_2\subseteq {\bf P}(y-y_{\alpha,Q_2},H).\] By algebraic manipulations, \[Q_1\subseteq {\bf P}(x_\alpha-y_{\alpha,Q_1},H),\ Q_1\subseteq{\bf P}(y-y_{\alpha, Q_1},H)\quad\Rightarrow\quad Q_1\subseteq{\bf P}(x_\alpha-y,H);\] similarly for $Q_2$. So $Q_1\cup Q_2\subseteq{\bf P}(x_\alpha-y,H)$ and hence $Q_1\cup Q_2\in I_\alpha$.] \begin{description} \item[$\otimes_3$] If $\bar Q=\langle Q_n:n\in\Gamma\rangle$ are pairwise disjoint subsets of ${\bf P}^*$, for some infinite $\Gamma\subseteq\omega$, then for some $n\in\Gamma$ we have $Q_n\in I_{\eta_\delta(n)}$. \end{description} [Why? Otherwise let $a_n$ be $Q_n$ if $n\in \Gamma$, and $\emptyset$ if $n\in \omega\setminus \Gamma$, and let $\bar a=\langle a_n: n< \omega\rangle$. Now $n\in\Gamma\quad\Rightarrow\quad\eta_\delta(n)\in {\rm inv}(y^{\delta, 0}_{\bar a},C_\delta;\bar H)$ and hence \[A_\delta\cap{\rm inv}(y^{\delta, 0}_{\bar a},C_\delta;\bar H)\supseteq\{\eta_\delta(n):n\in \Gamma\},\] which is infinite, contradicting the choice of $A_\delta$.] \begin{description} \item[$\otimes_4$] for all but finitely many $n$ the Boolean algebra ${\cal P}({\bf P}^*)/I_{\eta_\delta(n)}$ is finite. \end{description} [Why? If not, then by $\otimes_1$ second phrase, for each $n$ there are infinitely many non-principal ultrafilters $D$ on ${\bf P}^*$ disjoint to $I_{\eta_\delta(n)}$, so for $n<\omega$ we can find an ultrafilter $D_n$ on ${\bf P}^*$ disjoint to $I_{\eta_\delta(n)}$, distinct from $D_m$ for $m<n$. Thus we can find $\Gamma\in [\omega]^{\aleph_0}$ and $Q_n\in D_n$ for $n\in\Gamma$ such that $\langle Q_n:n\in\Gamma\rangle$ are pairwise disjoint (as $Q_n\in D_n$ clearly $|Q_n|=\aleph_0$). Why? Look: if $B_n \in D_0\setminus D_1$ for $n\in\omega$ then \[(\exists^\infty n)(B_n \in D_n)\quad\mbox{ or }\quad(\exists^\infty n) ({\bf P}^*\setminus B_n \in D_n),\] etc. Let $Q_n=\emptyset$ for $n\in\omega\setminus\Gamma$, now $\bar Q= \langle Q_n:n<\omega\rangle$ contradicts $\otimes_3$.] \begin{description} \item[$\otimes_5$] If the conclusion (of \ref{3.9}) fails, then for no $\alpha\in A_\delta$ is ${\cal P}({\bf P}^*)/I_\alpha$ finite. \end{description} [Why? If not, choose such an $\alpha$ and $Q^*\subseteq{\bf P}^*$, $Q^* \notin I_\alpha$ such that $I=I_\alpha\restriction Q^*$ is a maximal ideal on $Q^*$. So $D=:{\cal P}(Q^*)\setminus I$ is a non-principal ultrafilter. Remember $\beta=\beta_\alpha<\alpha$ is such that $\{y_{\alpha,Q}:Q\in I_\alpha\}\subseteq H_\beta$. Now, $H_\beta\in {\frak A}_{\beta+1}$, $|H_\beta|<\lambda$. Hence $(*)^2_\lambda$ from \ref{3.3} (note that it does not matter whether we consider an ordinal $\chi<\lambda$ or a cardinal $\chi<\lambda$, or any other set of cardinality $< \lambda$) implies that there is $S_{H_\beta}\in {\frak A}_{\beta+1}$, $S_{H_\beta}\subseteq [H_\beta]^{\le \aleph_0}$, $|S_{H_\beta}|<\lambda$ as there. Now it does not matter if we deal with functions from an ultrafilter on $\omega$ \underbar{or} an ultrafilter on $Q^*$. We define $f:D\longrightarrow H_\beta$ as follows: for $U\in D$ we let $f(U)=y_{\alpha,Q^* \setminus U}$. (Note: $Q^*\setminus U\in I_\alpha$, hence $y_{\alpha,Q^* \setminus U}$ is well defined.) So, by the choice of $S_{H_\beta}$ (see (ii) of $(*)^2_\lambda$), for some countable $f' \subseteq f$, $f'\in {\frak A}_{\beta+1}$ and $\bigcap\{U:U\in{\rm Dom}(f')\} \notin D$ (reflect for a minute). Let ${\rm Dom}(f')=\{U_0,U_1,\ldots\}$. Then $\bigcup\limits_{n<\omega}(Q^*\setminus U_n)\notin I_\alpha$. But as in the proof of $\otimes_2$, as \[\langle y_\alpha,(Q^* \setminus U_n):n<\omega\rangle\in {\frak A}_{\beta+1}\subseteq {\frak A}_\alpha,\] we have $\bigcup\limits_{n<\omega}(Q^*\setminus U_n)\in I_\alpha$, an easy contradiction.] Now $\otimes_4$, $\otimes_5$ give a contradiction. \hfill$\square_{\ref{3.3}}$ \begin{Remark} \label{3.10} We can deal similarly with $R$-modules, $|R|<\mu$ \underbar{if} $R$ has infinitely many prime ideals $I$. Also the treatment of ${\frak K}^{rs(p)}_\lambda$ is similar to the one for modules over rings with one prime. \noindent Note: if we replace ``reduced" by \[x\in G \setminus\{0\}\quad \Rightarrow\quad (\exists p\in{\bf P}^*)(x\notin pG)\] then here we could have defined \[{\bf P}(x,H)=:\{p\in {\bf P}^*:x\in pH\}\] and the proof would go through with no difference (e.g. choose a fixed partition $\langle {\bf P}^*_n: n< \omega\rangle$ of ${\bf P}^*$ to infinite sets, and let ${\bf P}'(x, H)=\{n: x\in pH\mbox{ for every }p\in {\bf P}^*_n\}$). Now the groups are less divisible. \end{Remark} \begin{Remark} \label{3.11} We can get that the groups are slender, in fact, the construction gives it. \end{Remark} \section{Below the continuum there may be universal structures} Both in \cite{Sh:456} (where we deal with universality for $(<\lambda)$-stable (Abelian) groups, like ${\frak K}^{rs(p)}_\lambda$) and in \S3, we restrict ourselves to $\lambda>2^{\aleph_0}$, a restriction which does not appear in \cite{KjSh:447}, \cite{KjSh:455}. Is this restriction necessary? In this section we shall show that at least to some extent, it is. We first show under MA that for $\lambda<2^{\aleph_0}$, any $G\in {\frak K}^{rs(p)}_\lambda$ can be embedded into a ``nice" one; our aim is to reduce the consistency of ``there is a universal in ${\frak K}^{rs(p)}_\lambda$" to ``there is a universal in ${\frak K}^{tr}_{\langle\aleph_0:n<\omega\rangle\char 94\langle \lambda \rangle}$". Then we proceed to prove the consistency of the latter. Actually a weak form of MA suffices. \begin{Definition} \label{4.2} \begin{enumerate} \item $G\in {\frak K}^{rs(p)}_\lambda$ is {\em tree-like} if: \begin{description} \item[(a)] we can find a basic subgroup $B= \bigoplus\limits_{\scriptstyle i<\lambda_n\atop\scriptstyle n<\omega} {\Bbb Z} x^n_i$, where \[\lambda_n=\lambda_n(G)=:\dim\left((p^nG)[p]/p^{n+1}(G)[p]\right)\] (see Fuchs \cite{Fu}) such that: ${\Bbb Z} x^n_i \cong {\Bbb Z}/p^{n+1} {\Bbb Z}$ and \begin{description} \item[$\otimes_0$] every $x\in G$ has the form \[\sum\limits_{n,i}\{a^n_i p^{n-k} x^n_i:n\in [k,\omega)\mbox{ and } i<\lambda\}\] where $a^n_i\in{\Bbb Z}$ and \[n<\omega\quad \Rightarrow\quad w_n[x]=:\{i:a^n_i\, p^{n-k}x^n_i\neq 0\} \mbox{ is finite}\] \end{description} (this applies to any $G\in {\frak K}^{rs(p)}_\lambda$ we considered so far; we write $w_n[x]=w_n[x,\bar Y]$ when $\bar Y=\langle x^n_i:n,i\rangle$). Moreover \item[(b)] $\bar Y=\langle x^n_i:n,i\rangle$ is tree-like inside $G$, which means that we can find $F_n:\lambda_{n+1}\longrightarrow\lambda_n$ such that letting $\bar F=\langle F_n:n<\omega\rangle$, $G$ is generated by some subset of $\Gamma(G,\bar Y,\bar F)$ where: \[\hspace{-0.5cm}\begin{ALIGN} \Gamma(G,\bar Y,\bar F)=\big\{x:&\mbox{for some }\eta\in\prod\limits_{n <\omega}\lambda_n, \mbox{ for each } n<\omega \mbox{ we have}\\ &F_n(\eta(n+1))=\eta(n)\mbox{ and }x=\sum\limits_{n\ge k} p^{n-k}x^n_{\eta(n)}\big\}. \end{ALIGN}\] \end{description} \item $G\in {\frak K}^{rs(p)}_\lambda$ is {\em semi-tree-like} if above we replace (b) by \begin{description} \item[(b)$'$] we can find a set $\Gamma\subseteq\{\eta:\eta$ is a partial function from $\omega$ to $\sup\limits_{n<\omega} \lambda_n$ with $\eta(n)< \lambda_n\}$ such that: \begin{description} \item[($\alpha$)] $\eta_1\in\Gamma,\ \eta_2\in\Gamma,\ \eta_1(n)=\eta_2(n) \quad \Rightarrow\quad\eta_1\restriction n=\eta_2\restriction n$, \item[($\beta$)] for $\eta\in\Gamma$ and $n\in{\rm Dom}(\eta)$, there is \[y_{\eta,n}=\sum \{p^{m-n}x^m_{\eta(m)}:m \in {\rm Dom}(\eta)\mbox{ and } m \ge n\}\in G,\] \item[($\gamma$)] $G$ is generated by \[\{x^n_i:n<\omega,i<\lambda_n\}\cup\{y_{\eta,n}:\eta\in\Gamma,n\in {\rm Dom}(\eta)\}.\] \end{description} \end{description} \item $G\in {\frak K}^{rs(p)}_\lambda$ is {\em almost tree-like} if in (b)$'$ we add \begin{description} \item[($\delta$)] for some $A\subseteq\omega$ for every $\eta\in\Gamma$, ${\rm Dom}(\eta)=A$. \end{description} \end{enumerate} \end{Definition} \begin{Proposition} \label{4.3} \begin{enumerate} \item Suppose $G\in {\frak K}^{rs(p)}_\lambda$ is almost tree-like, as witnessed by $A\subseteq\omega$, $\lambda_n$ (for $n<\omega$), $x^n_i$ (for $n\in A$, $i<\lambda_n$), and if $n_0<n_2$ are successive members of $A$, $n_0<n<n_2$ then $\lambda_n\ge\lambda_{n_0}$ or just \[\lambda_n\ge|\{\eta(n_0):\eta\in\Gamma\}|.\] {\em Then} $G$ is tree-like (possibly with other witnesses). \item If in \ref{4.2}(3) we just demand $\eta\in\Gamma\quad\Rightarrow \quad\bigvee\limits_{n<\omega}{\rm Dom}(\eta)\setminus n=A\setminus n$; then changing the $\eta$'s and the $y_{\eta,n}$'s we can regain the ``almost tree-like". \end{enumerate} \end{Proposition} \proof 1) For every successive members $n_0<n_2$ of $A$ for \[\alpha\in S_{n_0}=:\{\alpha:(\exists\eta)[\eta\in\Gamma\ \&\ \eta(n_0) =\alpha]\},\] choose ordinals $\gamma(n_0,\alpha,\ell)$ for $\ell\in (n_0,n_2)$ such that \[\gamma(n_0,\alpha_1,\ell)=\gamma(n_0,\alpha_2,\ell)\quad\Rightarrow \quad\alpha_1=\alpha_2.\] We change the basis by replacing for $\alpha\in S_{n_0}$, $\{x^n_\alpha\}\cup \{x^\ell_{\gamma(n_0,\alpha,\ell)}:\ell\in (n_0,n_2)\}$ (note: $n_0<n_2$ but possibly $n_0+1=n_2$), by: \[\begin{ALIGN} \biggl\{ x^{n_0}_\alpha + px^{n_0+1}_{\gamma(n_0,\alpha,n_0+1)}, &x^{n_0+1}_{\gamma(n_0,\alpha,n_0+1)} + px^{n_0+2}_{\gamma(n_0,\alpha,n_0+2)},\ldots, \\ &x^{n_2-2}_{\gamma(n_0,\alpha,n_2-2)} + px^{n_2-1}_{\gamma(n_0,\alpha,n_2-1)}, x^{n_2-1}_{\gamma(n_0,\gamma,n_2-1)} \biggr\}. \end{ALIGN}\] 2) For $\eta\in \Gamma$ let $n(\eta)=\min\{ n: n\in A\cap{\rm Dom}(\eta)$ and ${\rm Dom}(\eta)\setminus n=A\setminus n\}$, and let $\Gamma_n=\{\eta\in \Gamma: n(\eta)=n\}$ for $n\in A$. We choose by induction on $n< \omega$ the objects $\nu_\eta$ for $\eta\in \Gamma_n$ and $\rho^n_\alpha$ for $\alpha< \lambda_n$ such that: $\nu_\eta$ is a function with domain $A$, $\nu_\eta\restriction (A\setminus n(\eta))=\eta\restriction (A\setminus n(\eta))$ and $\nu_\eta\restriction (A\cap n(\eta))= \rho^n_{\eta(n)}$, $\nu_\eta(n)< \lambda_n$ and $\rho^n_\alpha$ is a function with domain $A\cap n$, $\rho^n_\alpha(\ell)< \lambda_\ell$ and $\rho^n_\alpha \restriction (A\cap \ell) = \rho^\ell_{\rho^n_\alpha(\ell)}$ for $\ell\in A\cap n$. There are no problems and $\{\nu_\eta: \eta\in \Gamma_n\}$ is as required. \hfill$\square_{\ref{4.3}}$ \begin{Theorem}[MA] \label{4.1} Let $\lambda<2^{\aleph_0}$. Any $G\in {\frak K}^{rs(p)}_\lambda$ can be embedded into some $G'\in {\frak K}^{rs(p)}_\lambda$ with countable density which is tree-like. \end{Theorem} \proof By \ref{4.3} it suffices to get $G'$ ``almost tree-like" and $A\subseteq\omega$ which satisfies \ref{4.3}(1). The ability to make $A$ thin helps in proving Fact E below. By \ref{1.1} without loss of generality $G$ has a base (i.e. a dense subgroup of the form) $B=\bigoplus\limits_{\scriptstyle n<\omega\atop\scriptstyle i<\lambda_n} {\Bbb Z} x^n_i$, where ${\Bbb Z} x^n_i\cong{\Bbb Z}/p^{n+1}{\Bbb Z}$ and $\lambda_n=\aleph_0$ (in fact $\lambda_n$ can be $g(n)$ if $g\in {}^\omega\omega$ is not bounded (by algebraic manipulations), this will be useful if we consider the forcing from \cite[\S2]{Sh:326}). Let $B^+$ be the extension of $B$ by $y^{n,k}_i$ ($k<\omega$, $n<\omega$, $i<\lambda_n$) generated freely except for $py^{n,k+1}_i=y^{n,k}_i$ (for $k<\omega$), $y^{n,\ell}_i=p^{n-\ell}x^n_i$ for $\ell\le n$, $n<\omega$, $i<\lambda_n$. So $B^+$ is a divisible $p$-group, let $G^+ =: B^+\bigoplus\limits_B G$. Let $\{z^0_\alpha:\alpha<\lambda\}\subseteq G[p]$ be a basis of $G[p]$ over $\{p^n x^n_i:n,i<\omega\}$ (as a vector space over ${\Bbb Z}/p{\Bbb Z}$ i.e. the two sets are disjoint, their union is a basis); remember $G[p]=\{x\in G:px=0\}$. So we can find $z^k_\alpha\in G$ (for $\alpha< \lambda$, $k<\omega$ and $k\ne 0$) such that \[pz^{k+1}_\alpha-z^k_\alpha=\sum_{i\in w(\alpha,k)} a^{k,\alpha}_i x^k_i,\] where $w(\alpha,k)\subseteq\omega$ is finite (reflect on the Abelian group theory). We define a forcing notion $P$ as follows: a condition $p \in P$ consists of (in brackets are explanations of intentions): \begin{description} \item[(a)] $m<\omega$, $M\subseteq m$, \end{description} [$M$ is intended as $A\cap\{0,\ldots,m-1\}$] \begin{description} \item[(b)] a finite $u\subseteq m\times\omega$ and $h:u\longrightarrow \omega$ such that $h(n,i)\ge n$, \end{description} [our extensions will not be pure, but still we want that the group produced will be reduced, now we add some $y^{n,k}_i$'s and $h$ tells us how many] \begin{description} \item[(c)] a subgroup $K$ of $B^+$: \[K=\langle y^{n,k}_i:(n,i)\in u,k<h(n,i)\rangle_{B^+},\] \item[(d)] a finite $w\subseteq\lambda$, \end{description} [$w$ is the set of $\alpha<\lambda$ on which we give information] \begin{description} \item[(e)] $g:w\rightarrow m + 1$, \end{description} [$g(\alpha)$ is in what level $m'\le m$ we ``start to think" about $\alpha$] \begin{description} \item[(f)] $\bar\eta=\langle\eta_\alpha:\alpha\in w\rangle$ (see (i)), \end{description} [of course, $\eta_\alpha$ is the intended $\eta_\alpha$ restricted to $m$ and the set of all $\eta_\alpha$ forms the intended $\Gamma$] \begin{description} \item[(g)] a finite $v\subseteq m\times\omega$, \end{description} [this approximates the set of indices of the new basis] \begin{description} \item[(h)] $\bar t=\{t_{n,i}:(n,i)\in v\}$ (see (j)), \end{description} [approximates the new basis] \begin{description} \item[(i)] $\eta_\alpha\in {}^M\omega$, $\bigwedge\limits_{\alpha\in w} \bigwedge\limits_{n\in M} (n,\eta_\alpha(n))\in v$, \end{description} [toward guaranteeing clause $(\delta)$ of \ref{4.2}(3) (see \ref{4.3}(2))] \begin{description} \item[(j)] $t_{n,i}\in K$ and ${\Bbb Z} t_{n,i} \cong {\Bbb Z}/p^n {\Bbb Z}$, \item[(k)] $K=\bigoplus\limits_{(n,i)\in v} ({\Bbb Z} t_{n,i})$, \end{description} [so $K$ is an approximation to the new basic subgroup] \begin{description} \item[(l)] if $\alpha\in w$, $g(\alpha)\le\ell\le m$ and $\ell\in M$ then \[z^\ell_\alpha-\sum\{t^{n-\ell}_{n,\eta_\alpha(n)}:\ell\le n\in {\rm Dom}(\eta_\alpha)\}\in p^{m-\ell}(K+G),\] \end{description} [this is a step toward guaranteeing that the full difference (when ${\rm Dom}(\eta_\alpha)$ is possibly infinite) will be in the closure of $\bigoplus\limits_{\scriptstyle n\in [i,\omega)\atop\scriptstyle i<\omega} {\Bbb Z} x^n_i$]. We define the order by: \noindent $p \le q$ \qquad if and only if \begin{description} \item[$(\alpha)$] $m^p\le m^q$, $M^q \cap m^p = M^p$, \item[$(\beta)$] $u^p\subseteq u^q$, $h^p\subseteq h^q$, \item[$(\gamma)$] $K^p\subseteq_{pr} K^q$, \item[$(\delta)$] $w^p\subseteq w^q$, \item[$(\varepsilon)$] $g^p\subseteq g^q$, \item[$(\zeta)$] $\eta^p_\alpha\trianglelefteq\eta^q_\alpha$, (i.e. $\eta^p_\alpha$ is an initial segment of $\eta^q_\alpha$) \item[$(\eta)$] $v^p\subseteq v^q$, \item[$(\theta)$] $t^p_{n,i}=t^q_{n,i}$ for $(n,i)\in v^p$. \end{description} \medskip \noindent{\bf A Fact}\hspace{0.15in} $(P,\le)$ is a partial order. \medskip \noindent{\em Proof of the Fact:}\ \ \ Trivial. \medskip \noindent{\bf B Fact}\hspace{0.15in} $P$ satisfies the c.c.c. (even is $\sigma$-centered). \medskip \noindent{\em Proof of the Fact:}\ \ \ It suffices to observe the following. Suppose that \begin{description} \item[$(*)$(i)] $p,q \in P$, \item[\quad(ii)] $M^p=M^q$, $m^p=m^q$, $h^p=h^q$, $u^p=u^q$, $K^p=K^q$, $v^p=v^q$, $t^p_{n,i}=t^q_{n,i}$, \item[\quad(iii)] $\langle\eta^p_\alpha:\alpha\in w^p\cap w^q\rangle = \langle\eta^q_\alpha:\alpha\in w^p\cap w^q\rangle$, \item[\quad(iv)] $g^p\restriction (w^p \cap w^q)=g^q \restriction(w^p \cap w^q)$. \end{description} Then the conditions $p,q$ are compatible (in fact have an upper bound with the same common parts): take the common values (in (ii)) or the union (for (iii)). \medskip \noindent{\bf C Fact}\hspace{0.15in} For each $\alpha<\lambda$ the set ${\cal I}_\alpha=:\{p\in P:\alpha\in w^p\}$ is dense (and open). \medskip \noindent{\em Proof of the Fact:}\ \ \ For $p\in P$ let $q$ be like $p$ except that: \[w^q=w^p\cup\{\alpha\}\quad\mbox{ and }\quad g^q(\beta)=\left\{ \begin{array}{lll} g^p(\beta) &\mbox{if}& \beta\in w^p \\ m^p &\mbox{if}& \beta=\alpha,\ \beta\notin w^p. \end{array}\right.\] \medskip \noindent{\bf D Fact}\hspace{0.15in} For $n<\omega$, $i<\omega$ the following set is a dense subset of $P$: \[{\cal J}^*_{(n,i)}=\{p\in P:x^n_i\in K^p\ \&\ (\forall n<m^p)(\{n\}\times \omega)\cap u^p \mbox{ has }>m^p\mbox{ elements}\}.\] \medskip \noindent{\em Proof of the Fact:}\ \ \ Should be clear. \medskip \noindent{\bf E Fact}\hspace{0.15in} For each $m<\omega$ the set ${\cal J}_m =:\{p\in P:m^p\ge m\}$ is dense in $P$. \medskip \noindent{\em Proof of the Fact:}\ \ \ Let $p\in P$ be given such that $m^p <m$. Let $w^p=\{\alpha_0,\ldots,\alpha_{r-1}\}$ be without repetitions; we know that in $G$, $pz^0_{\alpha_\ell}=0$ and $\{z^0_{\alpha_\ell}: \ell<r\}$ is independent $\mod\ B$, hence also in $K+G$ the set $\{z^0_{\alpha_\ell}:\ell<r\}$ is independent $\mod\ K$. Clearly \begin{description} \item[(A)] $pz^{k+1}_{\alpha_\ell}=z^k_{\alpha_\ell}\mod\ K$ for $k\in [g(\alpha_\ell),m^p)$, hence \item[(B)] $p^{m^p}z^{m^p}_{\alpha_\ell}=z^{g(\alpha_\ell)}_{\alpha_\ell} \mod\ K$. \end{description} Remember \begin{description} \item[(C)] $z^{m^p}_{\alpha_\ell}=\sum\{a^{k,\alpha_\ell}_i p^{k-m^p} x^k_i: k \ge m^p,i\in w(\alpha_\ell,k)\}$, \end{description} and so, in particular, (from the choice of $z^0_{\alpha_\ell}$) \[p^{m^p+1}z^{m^p}_{\alpha_\ell}=0\quad\mbox{ and }\quad p^{m^p}z^{m^p}_{\alpha_\ell}\ne 0.\] For $\ell<r$ and $n\in [m^p,\omega)$ let \[s^n_\ell=:\sum\big\{a^{k,\alpha_\ell}_i p^{k-m^p} x^k_i:k \ge m^p \mbox{ but }k<n\mbox{ and } i\in w(\alpha_\ell,k)\big\}.\] But $p^{k-m^p}x^k_i = y^{k,m^p}_i$, so \[s^n_\ell=\sum\big\{a^{k,\alpha_\ell}_i y^{k,m^p}_i:k\in [m^p,n) \mbox{ and }i\in (\alpha_\ell,k)\big\}.\] Hence, for some $m^*>m,m^p$ we have: $\{p^m\,s^{m^*}_\ell:\ell<r\}$ is independent in $G[p]$ over $K[p]$ and therefore in $\langle x^k_i:k\in [m^p,m^*],i<\omega\rangle$. Let \[s^*_\ell=\sum\big\{a^{k,\alpha_\ell}_i:k\in [m^p,m^*)\mbox{ and }i\in w(\alpha_\ell,k)\}.\] Then $\{ s^*_\ell:\ell<r\}$ is independent in \[B^+_{[m,m^*)}=\langle y^{l,m^*-1}_i:k\in [m^p,m^*)\mbox{ and }i<\omega \rangle.\] Let $i^*<\omega$ be such that: $w(\alpha_\ell,k)\subseteq\{0,\ldots,i^*-1\}$ for $k\in [m^p,m^*)$, $\ell=1,\ldots,r$. Let us start to define $q$: \[\begin{array}{c} m^q=m^*,\quad M^q=M^p\cup\{m^*-1\},\quad w^q=w^p,\quad g^q=g^p,\\ u^q=u^p\cup ([m^p,m^*)\times\{0,\ldots,i^*-1\}),\\ h^q\mbox{ is } h^p\mbox{ on } u^p\mbox{ and }h^q(k,i)=m^*-1\mbox{ otherwise},\\ K^q\mbox{ is defined appropriately, let } K'=\langle x^n_i:n\in [m^p,m^*), i<i^*\rangle. \end{array}\] Complete $\{s^*_\ell:\ell<r\}$ to $\{s^*_\ell:\ell<r^*\}$, a basis of $K'[p]$, and choose $\{t_{n,i}:(n,i)\in v^*\}$ such that: $[p^mt_{n,i}=0\ \ \Leftrightarrow\ \ m>n]$, and for $\ell<r$ \[p^{m^*-1-\ell}t_{m^*-1,\ell} = s^*_\ell.\] The rest should be clear. \medskip The generic gives a variant of the desired result: almost tree-like basis; the restriction to $M$ and $g$ but by \ref{4.3} we can finish. \hfill$\square_{\ref{4.2}}$ \begin{Conclusion} [MA$_\lambda$($\sigma$-centered)] \label{4.4} For $(*)_0$ to hold it suffices that $(*)_1$ holds where \begin{description} \item[$(*)_0$] in ${\frak K}^{rs(p)}_\lambda$, there is a universal member, \item[$(*)_1$] in ${\frak K}^{tr}_{\bar\lambda}$ there is a universal member, where: \begin{description} \item[(a)] $\lambda_n=\aleph_0$, $\lambda_\omega=\lambda$, $\ell g(\bar\lambda) =\omega+1$\qquad \underbar{or} \item[(b)] $\lambda_\omega=\lambda$, $\lambda_n\in [n,\omega)$, $\ell g (\bar \lambda)=\omega+1$. \end{description} \end{description} \end{Conclusion} \begin{Remark} \label{4.4A} Any $\langle\lambda_n:n<\omega\rangle$, $\lambda_n<\omega$ which is not bounded suffices. \end{Remark} \proof For case (a) - by \ref{4.1}. \noindent For case (b) - the same proof. \hfill$\square_{\ref{4.4}}$ \begin{Theorem} \label{4.5} Assume $\lambda<2^{\aleph_0}$ and \begin{description} \item[(a)] there are $A_i\subseteq\lambda$, $|A_i|=\lambda$ for $i<2^\lambda$ such that $i\ne j \Rightarrow |A_i \cap A_j| \le \aleph_0$. \end{description} Let $\bar\lambda=\langle \lambda_\alpha:\alpha\le\omega\rangle$, $\lambda_n = \aleph_0$, $\lambda_\omega=\lambda$. \noindent{\em Then} there is $P$ such that: \medskip \begin{description} \item[$(\alpha)$] $P$ is a c.c.c. forcing notion, \item[$(\beta)$] $|P|=2^\lambda$, \item[$(\gamma)$] in $V^P$, there is $T\in {\frak K}^{tr}_{\bar\lambda}$ into which every $T' \in ({\frak K}^{tr}_{\bar \lambda})^V$ can be embedded. \end{description} \end{Theorem} \proof Let $\bar T=\langle T_i:i<2^\lambda\rangle$ list the trees $T$ of cardinality $\le\lambda$ satisfying \[{}^{\omega >}\omega\subseteq T \subseteq {}^{\omega \ge} \omega\quad \mbox{ and }\quad T\cap {}^\omega\omega\mbox{ has cardinality $\lambda$, for simplicity.}\] Let $T_i\cap {}^\omega\omega=\{\eta^i_\alpha:\alpha\in A_i \}$. We shall force $\rho_{\alpha,\ell}\in {}^\omega\omega$ for $\alpha< \lambda$, $\ell<\omega$, and for each $i<2^\lambda$ a function $g_i:A_i \longrightarrow\omega$ such that: there is an automorphism $f_i$ of $({}^{\omega>}\omega,\triangleleft)$ which induces an embedding of $T_i$ into $\left(({}^{\omega>}\omega)\cup \{\rho_{\alpha,g_i(\alpha)}:\alpha< \lambda\},\triangleleft\right)$. We shall define $p\in P$ as an approximation. \noindent A condition $p\in P$ consists of: \begin{description} \item[(a)] $m<\omega$ and a finite subset $u$ of ${}^{m \ge}\omega$, closed under initial segments such that $\langle\rangle\in u$, \item[(b)] a finite $w\subseteq 2^\lambda$, \item[(c)] for each $i\in w$, a finite function $g_i$ from $A_i$ to $\omega$, \item[(d)] for each $i\in w$, an automorphism $f_i$ of $(u,\triangleleft)$, \item[(e)] a finite $v\subseteq\lambda\times\omega$, \item[(f)] for $(\alpha,n)\in v$, $\rho_{\alpha,n}\in u\cap ({}^m\omega)$, \end{description} such that \begin{description} \item[(g)] if $i\in w$ and $\alpha\in{\rm Dom}(g_i)$ then: \begin{description} \item[$(\alpha)$] $(\alpha,g_i(\alpha))\in v$, \item[$(\beta)$] $\eta^i_\alpha\restriction m\in u$, \item[$(\gamma)$] $f_i(\eta^i_\alpha\restriction m)=\rho_{\alpha, g_i(\alpha)}$, \end{description} \item[(h)] $\langle\rho_{\alpha,n}:(\alpha,n)\in v\rangle$ is with no repetition (all of length $m$), \item[(i)] for $i\in w$, $\langle\eta^i_\alpha\restriction m:\alpha\in {\rm Dom}(g_i)\rangle$ is with no repetition. \end{description} The order on $P$ is: $p \le q$ if and only if: \begin{description} \item[$(\alpha)$] $u^p \subseteq u^q$, $m^p\le m^q$, \item[$(\beta)$] $w^p \subseteq w^q$, \item[$(\gamma)$] $f^p_i \subseteq f^q_i$ for $i\in w^p$, \item[$(\delta)$] $g^p_i \subseteq g^q_i$ for $i\in w^p$, \item[$(\varepsilon)$] $v^p \subseteq v^q$, \item[$(\zeta)$] $\rho^p_{\alpha,n}\trianglelefteq\rho^q_{\alpha,n}$, when $(\alpha,n) \in v^p$, \item[$(\eta)$] if $i\ne j\in w^p$ then for every $\alpha\in A_i\cap A_j\setminus ({\rm Dom}(g^p_i)\cap {\rm Dom}(g^p_j))$ we have $g^q_i(\alpha)\ne g^q_j(\alpha)$. \end{description} \medskip \noindent{\bf A Fact}\hspace{0.15in} $(P,\le)$ is a partial order. \medskip \noindent{\em Proof of the Fact:}\ \ \ Trivial. \medskip \noindent{\bf B Fact}\hspace{0.15in} For $i<2^\lambda$ the set $\{p:i\in w^p\}$ is dense in $P$. \medskip \noindent{\em Proof of the Fact:}\ \ \ If $p\in P$, $i\in 2^\lambda \setminus w^p$, define $q$ like $p$ except $w^q=w^p\cup\{i\}$, ${\rm Dom}(g^q_i)=\emptyset$. \medskip \noindent{\bf C Fact}\hspace{0.15in} If $p\in P$, $m_1\in (m^p,\omega)$, $\eta^*\in u^p$, $m^*<\omega$, $i\in w^p$, $\alpha\in\lambda \setminus{\rm Dom}(g^p_i)$ {\em then} we can find $q$ such that $p\le q\in P$, $m^q>m_1$, $\eta^* \char 94\langle m^*\rangle\in u^q$ and $\alpha\in {\rm Dom}(g_i)$ and $\langle\eta^j_\beta\restriction m^q:j\in w^q$ and $\beta\in {\rm Dom}(g^q_j)\rangle$ is with no repetition, more exactly $\eta^{j(1)}_{\beta_1}\setminus m^q= \eta^{j(2)}_{\beta_2}\restriction m^q \Rightarrow \eta^{j(1)}_{\beta_1}=\eta^{j(2)}_{\beta_2}$. \medskip \noindent{\em Proof of the Fact:}\ \ \ Let $n_0\le m^p$ be maximal such that $\eta^i_\alpha\restriction n_0 \in u^p$. Let $n_1<\omega$ be minimal such that $\eta^i_\alpha\restriction n_1\notin\{\eta^i_\beta\restriction n_1: \beta\in{\rm Dom}(g^p_i)\}$ and moreover the sequence \[\langle\eta^j_\beta\restriction n_1:j\in w^p\ \&\ \beta\in{\rm Dom}(g^p_j)\ \ \mbox{ or }\ \ j=i\ \&\ \beta=\alpha\rangle\] is with no repetition. Choose a natural number $m^q>m^p+1,n_0+1,n_1+2$ and let $k^*=:3+\sum\limits_{i\in w^p}|{\rm Dom}(g^p_i)|$. Choose $u^q\subseteq {}^{m^q\ge}\omega$ such that: \begin{description} \item[(i)] $u^p\subseteq u^q\subseteq {}^{m^q\ge}\omega$, $u^q$ is downward closed, \item[(ii)] for every $\eta\in u^q$ such that $\ell g(\eta)<m^q$, for exactly $k^*$ numbers $k$, $\eta\char 94\langle k\rangle\in u^q\setminus u^p$, \item[(iii)] $\eta^j_\beta\restriction\ell\in u^q$ when $\ell\le m^q$ and $j\in w^p$, $\beta\in{\rm Dom}(g^p_j)$, \item[(iv)] $\eta^i_\alpha\restriction\ell\in u^q$ for $\ell\le m^q$, \item[(v)] $\eta^*\char 94\langle m^* \rangle\in u^q$. \end{description} Next choose $\rho^q_{\beta,n}$ (for pairs $(\beta,n)\in v^p)$ such that: \[\rho^p_{\beta,n}\trianglelefteq\rho^q_{\beta,n}\in u^q\cap {}^{m^q} \omega.\] For each $j\in w^p$ separately extend $f^p_j$ to an automorphism $f^q_j$ of $(u^q,\triangleleft)$ such that for each $\beta\in{\rm Dom}(g^p_j)$ we have: \[f^q_j(\eta^j_\beta\restriction m^q)=\rho^q_{\beta,g_j}(\beta).\] This is possible, as for each $\nu\in u^p$, and $j\in w^p$, we can separately define \[f^q_j\restriction\{\nu':\nu\triangleleft\nu'\in u^q\ \mbox{ and }\ \nu' \restriction (\ell g(\nu)+1)\notin u^p\}\] --its range is \[\{\nu':f^p_j(\nu)\triangleleft \nu'\in u^q\ \mbox{ and }\ \nu' \restriction (\ell g(\nu)+1)\notin u^p\}.\] The point is: by Clause (ii) above those two sets are isomorphic and for each $\nu$ at most one $\rho^p_{\beta,n}$ is involved (see Clause (h) in the definition of $p \in P$). Next let $w^q=w^p$, $g^q_j=g^p_j$ for $j\in w \setminus\{i\}$, $g^q_i\restriction{\rm Dom}(g^p_i)=g^p_i$, $g^q_i(\alpha)= \min(\{n:(\alpha,n)\notin v^p\})$, ${\rm Dom}(g^q_i)={\rm Dom}(g^p_i)\cup\{\alpha\}$, and $\rho^q_{\alpha,g^q_i(\alpha)}=f^g_i(\eta^i_\alpha\restriction m^q)$ and $v^q=v^p\cup\{(\alpha,g^q_i(\alpha))\}$. \medskip \noindent{\bf D Fact}\hspace{0.15in} $P$ satisfies the c.c.c. \medskip \noindent{\em Proof of the Fact:}\ \ \ Assume $p_\varepsilon\in P$ for $\varepsilon<\omega_1$. By Fact C, without loss of generality each \[\langle\eta^j_\beta\restriction m^{p_\varepsilon}:j\in w^{p_\varepsilon} \mbox{ and }\beta\in{\rm Dom}(g^{p_\varepsilon}_j)\rangle\] is with no repetition. Without loss of generality, for all $\varepsilon <\omega_1$ \[U_\varepsilon=:\big\{\alpha<2^\lambda:\alpha\in w^{p_\varepsilon}\mbox{ or } \bigvee_{i\in w^p}[\alpha\in{\rm Dom}(g_i)]\mbox{ or }\bigvee_k(k,\alpha)\in v^{p_\varepsilon}\big\}\] has the same number of elements and for $\varepsilon\ne\zeta<\omega_1$, there is a unique one-to-one order preserving function from $U_\varepsilon$ onto $U_\zeta$ which we call ${\rm OP}_{\zeta,\varepsilon}$, which also maps $p_\varepsilon$ to $p_\zeta$ (so $m^{p_\zeta}=m^{p_\varepsilon}$; $u^{p_\zeta} =u^{p_\varepsilon}$; ${\rm OP}_{\zeta,\varepsilon}(w^{p_\varepsilon})= w^{p_\zeta}$; if $i\in w^{p_\varepsilon}$, $j={\rm OP}_{\zeta,\varepsilon}(i)$, then $f_i\circ{\rm OP}_{\varepsilon,\zeta}\equiv f_j$; and {\em if\/} $\beta= {\rm OP}_{\zeta,\varepsilon}(\alpha)$ and $\ell<\omega$ {\em then} \[(\alpha,\ell)\in v^{p_\varepsilon}\quad\Leftrightarrow\quad (\beta,\ell) \in v^{p_\zeta}\quad\Rightarrow\quad\rho^{p_\varepsilon}_{\alpha,\ell}= \rho^{p_\zeta}_{\beta,\ell}).\] Also this mapping is the identity on $U_\zeta\cap U_\varepsilon$ and $\langle U_\zeta:\zeta<\omega_1\rangle$ is a $\triangle$-system. Let $w=:w^{p_0}\cap w^{p_1}$. As $i\ne j\ \Rightarrow\ |A_i\cap A_j|\le \aleph_0$, without loss of generality \begin{description} \item[$(*)$] if $i\ne j\in w$ then \[U_{\varepsilon}\cap (A_i\cap A_j)\subseteq w.\] \end{description} We now start to define $q\ge p_0,p_1$. Choose $m^q$ such that $m^q\in (m^{p_\varepsilon},\omega)$ and \[\begin{array}{ll} m^q>\max\big\{\ell g(\eta^{i_0}_{\alpha_0}\cap\eta^{i_1}_{\alpha_1})+1:& i_0\in w^{p_0},\ i_1 \in w^{p_1},\ {\rm OP}_{1,0}(i_0)=i_1,\\ \ &\alpha_0\in{\rm Dom}(g^{p_0}_{i_0}),\ \alpha_1\in{\rm Dom}(g^{p_1}_{i_1}),\\ \ &{\rm OP}_{1,0}(\alpha_0)=\alpha_1\big\}. \end{array}\] Let $u^q\subseteq {}^{m^q\ge}\omega$ be such that: \begin{description} \item[(A)] $u^q\cap\left({}^{m^{p_0}\ge}\omega\right)=u^q\cap\left( {}^{m^{p_1}\ge}\omega\right)=u^{p_0}=u^{p_1}$, \item[(B)] for each $\nu\in u^q$, $m^{p_0}\le\ell g(\nu)<m^q$, for exactly two numbers $k<\omega$, $\nu\char 94 \langle k\rangle\in u^q$, \item[(C)] $\eta^i_\alpha\restriction\ell\in u^q$ for $\ell\le m^q$ \underbar{when}: $i\in w^{p_0}$, $\alpha\in{\rm Dom}(g^{p_0}_i)$ \underbar{or} $i\in w^{p_1}$, $\alpha\in{\rm Dom}(g^{p_1}_i)$. \end{description} [Possible as $\{\eta^i_\alpha\restriction m^{p_\varepsilon}:i\in w^{p_\varepsilon},\alpha\in{\rm Dom}(g^{p_\varepsilon}_i)\}$ is with no repetitions (the first line of the proof).] Let $w^q=:w^{p_0}\cup w^{p_1}$ and $v^q=:v^{p_0}\cup v^{p_1}$ and for $i \in w^q$ \[g^q_i=\left\{\begin{array}{lll} g^{p_0}_i &\mbox{\underbar{if}}& i\in w^{p_0}\setminus w^{p_1},\\ g^{p_1}_i &\mbox{\underbar{if}}& i\in w^{p_1}\setminus w^{p_0},\\ g^{p_0}_i \cup g^{p_1}_i &\mbox{\underbar{if}}& i\in w^{p_0}\cap w^{p_1}. \end{array}\right.\] Next choose $\rho^q_{\alpha,\ell}$ for $(\alpha,\ell)\in v^q$ as follows. Let $\nu_{\alpha,\ell}$ be $\rho^{p_0}_{\alpha,\ell}$ if defined, $\rho^{p_1}_{\alpha,\ell}$ if defined (no contradiction). If $(\alpha,\ell) \in v^q$ choose $\rho^q_{\alpha,\ell}$ as any $\rho$ such that: \begin{description} \item[$\otimes_0$] $\nu_{\alpha,\ell}\triangleleft\rho\in u^q\cap {}^{(m^q)}\omega$. \end{description} But not all choices are O.K., as we need to be able to define $f^q_i$ for $i\in w^q$. A possible problem will arise only when $i\in w^{p_0}\cap w^{p_1}$. Specifically we need just (remember that $\langle \rho^{p_\varepsilon}_{\alpha,\ell}:(\alpha,\ell)\in v^{p_\varepsilon} \rangle$ are pairwise distinct by clause (b) of the Definition of $p\in P$): \begin{description} \item[$\otimes_1$] if $i_0\in w^{p_0},(\alpha_0,\ell)=(\alpha_0, g_{i_0}(\alpha_0)),\alpha_0\in{\rm Dom}(g^{p_0}_{i_0})$, $i_1={\rm OP}_{1,0}(i_0)$ and $\alpha_1={\rm OP}_{1,0}(\alpha_0)$ and $i_0=i_1$ {\em then} $\ell g(\eta^{i_0}_{\alpha_0}\cap\eta^{i_1}_{\alpha_1})=\ell g (\rho^q_{\alpha_0,\ell}\cap\rho^q_{\alpha_1,\ell})$. \end{description} We can, of course, demand $\alpha_0\neq \alpha_1$ (otherwise the conclusion of $\otimes_1$ is trivial). Our problem is expressible for each pair $(\alpha_0,\ell),(\alpha_1,\ell)$ separately as: first the problem is in defining the $\rho^q_{(\alpha,\ell)}$'s and second, if $(\alpha'_1,\ell')$, $(\alpha'_2,\ell)$ is another such pair then $\{(\alpha_1,\ell), (\alpha_2,\ell)\}$, $\{(\alpha'_1,\ell'),(\alpha'_2,\ell')\}$ are either disjoint or equal. Now for a given pair $(\alpha_0,\ell),(\alpha_1,\ell)$ how many $i_0=i_1$ do we have? Necessarily $i_0\in w^{p_0}\cap w^{p_1}=w$. But if $i'_0\ne i''_0$ are like that then $\alpha_0\in A_{i'_0}\cap A_{i''_0}$, contradicting $(*)$ above because $\alpha_0\neq \alpha_1={\rm OP}_{1,0}(\alpha_0)$. So there is at most one candidate $i_0=i_1$, so there is no problem to satisfy $\otimes_1$. Now we can define $f^q_i$ (i$\in w^q$) as in the proof of Fact C. \medskip The rest should be clear. \hfill$\square_{\ref{4.4}}$ \begin{Conclusion} \label{4.6} Suppose $V\models GCH$, $\aleph_0<\lambda<\chi$ and $\chi^\lambda=\chi$. Then for some c.c.c. forcing notion $P$ of cardinality $\chi$, not collapsing cardinals nor changing cofinalities, in $V^P$: \begin{description} \item[(i)] $2^{\aleph_0}=2^\lambda=\chi$, \item[(ii)] ${\frak K}^{tr}_\lambda$ has a universal family of cardinality $\lambda^+$, \item[(iii)] ${\frak K}^{rs(p)}_\lambda$ has a universal family of cardinality $\lambda^+$. \end{description} \end{Conclusion} \proof First use a preliminary forcing $Q^0$ of Baumgartner \cite{B}, adding $\langle A_\alpha:\alpha<\chi\rangle$, $A_\alpha\in [\lambda]^\lambda$, $\alpha\neq\beta\quad \Rightarrow\quad |A_\alpha\cap A_\beta|\le\aleph_0$ (we can have $2^{\aleph_0}=\aleph_1$ here, or $[\alpha\ne \beta\quad \Rightarrow\quad A_\alpha\cap A_\beta$ finite], but not both). Next use an FS iteration $\langle P_i,\dot{Q}_i:i<\chi\times \lambda^+\rangle$ such that each forcing from \ref{4.4} appears and each forcing as in \ref{4.5} appears. \hfill$\square_{\ref{4.6}}$ \begin{Remark} \label{4.7} We would like to have that there is a universal member in ${\frak K}^{rs(p)}_\lambda$; this sounds very reasonable but we did not try. In our framework, the present result shows limitations to ZFC results which the methods applied in the previous sections can give. \end{Remark} \section{Back to ${\frak K}^{rs(p)}$, real non-existence results} By \S1 we know that if $G$ is an Abelian group with set of elements $\lambda$, $C\subseteq\lambda$, then for an element $x\in G$ the distance from $\{y:y<\alpha\}$ for $\alpha\in C$ does not code an appropriate invariant. If we have infinitely many such distance functions, e.g. have infinitely many primes, we can use more complicated invariants related to $x$ as in \S3. But if we have one prime, this approach does not help. If one element fails, can we use infinitely many? A countable subset $X$ of $G$ can code a countable subset of $C$: \[\{\alpha\in C:\mbox{ closure}(\langle X\rangle_G)\cap\alpha\nsubseteq \sup(C\cap\alpha)\},\] but this seems silly - we use heavily the fact that $C$ has many countable subsets (in particular $>\lambda$) and $\lambda$ has at least as many. However, what if $C$ has a small family (say of cardinality $\le\lambda$ or $<\mu^{\aleph_0}$) of countable subsets such that every subset of cardinality, say continuum, contains one? Well, we need more: we catch a countable subset for which the invariant defined above is infinite (necessarily it is at most of cardinality $2^{\aleph_0}$, and because of \S4 we are not trying any more to deal with $\lambda\le 2^{\aleph_0}$). The set theory needed is expressed by $T_J$ below, and various ideals also defined below, and the result itself is \ref{5.7}. Of course, we can deal with other classes like torsion free reduced groups, as they have the characteristic non-structure property of unsuperstable first order theories; but the relevant ideals will vary: the parallel to $I^0_{\bar\mu}$ for $\bigwedge\limits_n \mu_n=\mu$, $J^2_{\bar\mu}$ seems to be always O.K. \begin{Definition} \label{5.1} \begin{enumerate} \item For $\bar\mu=\langle\mu_n:n<\omega\rangle$ let $B_{\bar\mu}$ be \[\bigoplus\{K^n_\alpha:n<\omega,\alpha<\mu_n\},\qquad K^n_\alpha= \langle{}^* t^n_\alpha\rangle_{K^n_\alpha}\cong {\Bbb Z}/p^{n+1} {\Bbb Z}.\] Let $B_{\bar\mu\restriction n}=\bigoplus\{K^m_\alpha:\alpha<\mu_m,m<n\} \subseteq B_{\bar\mu}$ (they are in ${\frak K}^{rs(p)}_{\le\sum\limits_{n} \mu_n}$). Let $\hat B$ be the $p$-torsion completion of $B$ (i.e. completion under the norm $\|x\|=\min\{2^{-n}:p^n\mbox{ divides }x\}$). \item Let $I^1_{\bar \mu}$ be the ideal on $\hat B_{\bar \mu}$ generated by $I^0_{\bar \mu}$, where \[\begin{array}{ll} I^0_{\bar\mu}=\big\{A\subseteq\hat B_{\bar\mu}:&\mbox{for every large enough }n,\\ &\mbox{for no }y\in\bigoplus\{K^m_\alpha:m\le n\mbox{ and }\alpha<\mu_m\}\\ &\mbox{but }y\notin\bigoplus\{K^m_\alpha:m<n\mbox{ and }\alpha<\mu_m\} \mbox{ we have}:\\ &\mbox{for every }m\mbox{ for some }z\in\langle A\rangle\mbox{ we have:}\\ &p^m\mbox{ divides }z-y\big\}. \end{array}\] (We may write $I^0_{\hat B_{\bar\mu}}$, but the ideal depends also on $\langle\bigoplus\limits_{\alpha<\mu_n} K^n_\alpha:n<\omega\rangle$ not just on $\hat B_{\bar\mu}$ itself). \item For $X,A\subseteq\hat B_{\bar\mu}$, \[\mbox{ recall }\ \ \langle A\rangle_{\bar B_{\bar\mu}}=\big\{\sum\limits_{n<n^*} a_ny_n:y_n \in A,\ a_n\in{\Bbb Z}\mbox{ and } n^*\in{\Bbb N}\big\},\] \[\mbox{ and let }\ \ c\ell_{\hat B_{\bar\mu}}(X)=\{x:(\forall n)(\exists y\in X)(x-y\in p^n \hat B_{\bar \mu})\}.\] \item Let $J^1_{\bar \mu}$ be the ideal which $J^{0.5}_{\bar \mu}$ generates, where \[\begin{array}{ll} J^{0.5}_{\bar\mu}=\big\{A\subseteq\prod\limits_{n<\omega}\mu_n: &\mbox{for some }n<\omega\mbox{ for no }m\in [n,\omega)\\ &\mbox{and }\beta<\gamma<\mu_m\mbox{ do we have}:\\ &\mbox{for every }k\in [m,\omega)\mbox{ there are }\eta,\nu\in A\mbox{ such}\\ &\mbox{that:}\ \ \eta(m)=\beta,\,\nu(m)=\gamma,\,\eta\restriction m= \nu\restriction m\\ &\mbox{and }\eta\restriction (m,k)=\nu\restriction (m,k)\big\}. \end{array}\] \item \[\begin{array}{rr} J^0_{\bar\mu}=\{A\subseteq\prod\limits_{n<\omega}\mu_n: &\mbox{for some }n<\omega\mbox{ and } k,\mbox{ the mapping }\eta\mapsto \eta \restriction n\\ &\mbox{is }(\le k)\mbox{-to-one }\}. \end{array}\] \item $J^2_{\bar\mu}$ is the ideal of nowhere dense subsets of $\prod\limits_n\mu_n$ (under the following natural topology: a neighbourhood of $\eta$ is $U_{\eta,n}=\{\nu:\nu\restriction n=\eta \restriction n\}$ for some $n$). \item $J^3_{\bar \mu}$ is the ideal of meagre subsets of $\prod\limits_n \mu_n$, i.e. subsets which are included in countable union of members of $J^2_{\bar \mu}$. \end{enumerate} \end{Definition} \begin{Observation} \label{5.2} \begin{enumerate} \item $I^0_{\bar\mu}$, $J^0_{\bar\mu}$, $J^{0.5}_{\bar\mu}$ are $(<\aleph_1)$-based, i.e. for $I^0_{\bar \mu}$: if $A\subseteq\hat B_{\bar\mu}$, $A\notin I^0_{\bar\mu}$ then there is a countable $A_0 \subseteq A$ such that $A_0\notin I^0_{\bar\mu}$. \item $I^1_{\bar\mu}$, $J^0_{\bar\mu}$, $J^1_{\bar\mu}$, $J^2_{\bar\mu}$, $J^3_{\bar\mu}$ are ideals, $J^3_{\bar\mu}$ is $\aleph_1$-complete. \item $J^0_{\bar\mu}\subseteq J^1_{\bar\mu}\subseteq J^2_{\bar\mu} \subseteq J^3_{\bar\mu}$. \item There is a function $g$ from $\prod\limits_{n<\omega}\mu_n$ into $\hat B_{\bar\mu}$ such that for every $X\subseteq\prod\limits_{n<\omega} \mu_n$: \[X\notin J^1_{\bar\mu}\quad \Rightarrow\quad g''(X)\notin I^1_{\bar\mu}.\] \end{enumerate} \end{Observation} \proof E.g. 4)\ \ Let $g(\eta)=\sum\limits_{n<\omega}p^n({}^*t^n_{\eta(n)})$. Let $X \subseteq\prod\limits_{n<\omega}\mu_n$, $X\notin J^1_{\bar\mu}$. Assume $g''(X)\in\bar I^1_{\bar\mu}$, so for some $\ell^*$ and $A_\ell \subseteq\hat B_{\bar\mu}$, ($\ell<\ell^*$) we have $A_\ell\in I^0_{\bar\mu}$, and $g''(X)\subseteq\bigcup\limits_{\ell<\ell^*} A_\ell$, so $X= \bigcup\limits_{\ell<\ell^*} X_\ell$, where \[X_\ell=:\{\eta\in X:g(\eta)\in A_\ell\}.\] As $J^1_{\bar \mu}$ is an ideal, for some $\ell<\ell^*$, $X_\ell\notin J^1_{\bar\mu}$. So by the definition of $J^1_{\bar\mu}$, for some infinite $\Gamma\subseteq\omega$ for each $m\in\Gamma$ we have $\beta_m<\gamma_m< \mu_m$ and for every $k\in [m,\omega)$ we have $\eta_{m,k},\nu_{m,k}$, as required in the definition of $J^1_{\bar \mu}$. So $g(\eta_{m,k}), g(\nu_{m,k}) \in A_\ell$ (for $m\in \Gamma$, $k\in (m,\omega)$). Now \[{}^* t^m_{\gamma_m} - {}^* t^m_{\beta_m}=g(\eta_{m,k})-g(\nu_{m,k}) \mod\ p^k \hat B_{\bar \mu},\] but $g(\eta_{m,k})-g(\nu_{m,k})\in\langle A_\ell\rangle_{\hat B_{\bar\mu}}$. Hence \[(\exists z\in\langle A_\ell\rangle_{\hat B_{\bar\mu}})[{}^* t^m_{\gamma_m} -{}^* t^m_{\beta_m}=z\mod\ p^k \hat B_{\bar\mu}],\] as this holds for each $k$, ${}^* t^m_{\gamma_m}-{}^* t^m_{\beta_m}\in c \ell(\langle A_\ell\rangle_{\hat B_{\bar\mu}})$. This contradicts $A_\ell \in I^0_{\bar \mu}$. \hfill$\square_{\ref{5.2}}$ \begin{Definition} \label{5.3} Let $I\subseteq{\cal P}(X)$ be downward closed (and for simplicity $\{\{x\}: x\in X\}\subseteq I$). Let $I^+={\cal P}(X)\setminus I$. Let \[\begin{array}{ll} {\bf U}^{<\kappa}_I(\mu)=:\min\big\{|{\cal P}|:&{\cal P}\subseteq [\mu]^{<\kappa}, \mbox{ and for every } f:X\longrightarrow\mu\mbox{ for some}\\ &Y\in {\cal P},\mbox{ we have }\{x\in X:f(x)\in Y\}\in I^+\big\}. \end{array}\] Instead of $<\kappa^+$ in the superscript of ${\bf U}$ we write $\kappa$. If $\kappa>|{\rm Dom}(I)|^+$, we omit it (since then its value does not matter). \end{Definition} \begin{Remark} \label{5.4} \begin{enumerate} \item If $2^{<\kappa}+|{\rm Dom}(I)|^{<\kappa}\le\mu$ we can find $F\subseteq$ partial functions from ${\rm Dom}(I)$ to $\mu$ such that: \begin{description} \item[(a)] $|F|={\bf U}^{<\kappa}_I(\mu)$, \item[(b)] $(\forall f:X\longrightarrow\mu)(\exists Y\in I^+)[f\restriction Y \in F]$. \end{description} \item Such functions (as ${\bf U}^{<\kappa}_I(\mu)$) are investigated in {\bf pcf} theory (\cite{Sh:g}, \cite[\S6]{Sh:410}, \cite[\S2]{Sh:430}, \cite{Sh:513}). \item If $I\subseteq J\subseteq {\cal P}(X)$, then ${\bf U}^{<\kappa}_I(\mu)\le {\bf U}^{<\kappa}_J(\mu)$, hence by \ref{5.2}(3), and the above \[{\bf U}^{<\kappa}_{J^0_{\bar\mu}}(\mu)\le {\bf U}^{<\kappa}_{J^1_{\bar\mu}}(\mu) \le {\bf U}^{<\kappa}_{J^2_{\bar\mu}}(\mu)\le {\bf U}^{<\kappa}_{J^3_{\bar\mu}}(\mu)\] and by \ref{5.2}(4) we have ${\bf U}^{<\kappa}_{I^1_{\bar \mu}}\leq {\bf U}^{<\kappa}_{J^1_{\bar \mu}}(\mu).$ \item On ${\rm IND}_\theta(\bar\kappa)$ (see \ref{5.5A} below) see \cite{Sh:513}. \end{enumerate} \end{Remark} \begin{Definition} \label{5.5A} ${\rm IND}'_\theta(\langle\kappa_n:n<\omega\rangle)$ means that for every model $M$ with universe $\bigcup\limits_{n<\omega}\kappa_n$ and $\le\theta$ functions, for some $\Gamma\in [\omega]^{\aleph_0}$ and $\eta\in \prod\limits_{n<\omega}\kappa_n$ we have: \[n\in\Gamma\quad\Rightarrow\quad\eta(n)\notin c\ell_M\{\eta(\ell):\ell \ne n\}.\] \end{Definition} \begin{Remark} Actually if $\theta\geq \aleph_0$, this implies that we can fix $\Gamma$, hence replacing $\langle \kappa_n: n< \omega\rangle$ by an infinite subsequence we can have $\Gamma=\omega$. \end{Remark} \begin{Theorem} \label{5.5} \begin{enumerate} \item If $\mu_n\rightarrow (\kappa_n)^2_{2^\theta}$ and ${\rm IND}'_\theta(\langle \kappa_n:n<\omega\rangle)$ {\em then} $\prod\limits_{n<\omega}\mu_n$ is not the union of $\le\theta$ sets from $J^1_{\bar \mu}$. \item If $\theta=\theta^{\aleph_0}$ and $\neg{\rm IND}'_\theta(\langle\mu_n: n<\omega\rangle$) then $\prod\limits_{n<\omega}\mu_n$ is the union of $\le\theta$ members of $J^1_{\bar\mu}$. \item If $\lim\sup\limits_n \mu_n$ is $\ge 2$, then $\prod\limits_{n<\omega} \mu_n\notin J^3_{\bar\mu}$ (so also the other ideals defined above are not trivial by \ref{5.2}(3), (4)). \end{enumerate} \end{Theorem} \proof 1)\ \ Suppose $\prod\limits_{n<\omega}\mu_n$ is $\bigcup\limits_{i<\theta} X_i$, and each $X_i\in J^1_{\bar\mu}$. We define for each $i<\theta$ and $n<k<\omega$ a two-place relation $R^{n,k}_i$ on $\mu_n$: \qquad $\beta R^{n,k}_i \gamma$ if and only if \qquad there are $\eta,\nu\in X_i\subseteq\prod\limits_{\ell<k}\mu_\ell$ such that \[\eta\restriction [0,n)=\nu\restriction [0,n)\quad\mbox{and }\ \eta\restriction (n,k)=\nu\restriction (n,k)\quad\mbox{and }\ \eta(n) =\beta,\ \nu(n)=\gamma.\] Note that $R^{n,k}_i$ is symmetric and \[n<k_1<k_2\ \&\ \beta R^{n,k_2}_i \gamma\quad \Rightarrow\quad \beta R^{n,k_1}_i \gamma.\] As $\mu_n\rightarrow (\kappa_n)^2_{2^\theta}$, we can find $A_n\in [\mu_n]^{\kappa_n}$ and a truth value ${\bf t}^{n,k}_i$ such that for all $\beta<\gamma$ from $A_n$, the truth value of $\beta R^{n,k}_i\gamma$ is ${\bf t}^{n,k}_i$. If for some $i$ the set \[\Gamma_i=:\{n<\omega:\mbox{ for every }k\in (n,\omega)\mbox{ we have } {\bf t}^{n,k}_i=\mbox{ true}\}\] is infinite, we get a contradiction to ``$X_i\in J^1_{\bar \mu}$", so for some $n(i)<\omega$ we have $n(i)=\sup(\Gamma_i)$. For each $n<k<\omega$ and $i<\theta$ we define a partial function $F^{n,k}_i$ from $\prod\limits_{\scriptstyle \ell<k,\atop\scriptstyle\ell \ne n} A_\ell$ into $A_n$: \begin{quotation} \noindent $F(\alpha_0\ldots\alpha_{n-1},\alpha_{n+1},\ldots,\alpha_k)$ is the first $\beta\in A_n$ such that for some $\eta\in X_i$ we have \[\begin{array}{c} \eta\restriction [0,n)=\langle\alpha_0,\ldots,\alpha_{n-1}\rangle,\quad \eta(n)=\beta,\\ \eta\restriction (n,k)=\langle\alpha_{n+1},\ldots,\alpha_{k-1}\rangle. \end{array}\] \end{quotation} So as ${\rm IND}'_\theta(\langle\kappa_n:n<\omega\rangle)$ there is $\eta=\langle \beta_n:n<\omega\rangle\in\prod\limits_{n<\omega} A_n$ such that for infinitely many $n$, $\beta_n$ is not in the closure of $\{\beta_\ell:\ell <\omega,\,\ell\ne n\}$ by the $F^{n,k}_i$'s. As $\eta\in\prod\limits_{n< \omega} A_n\subseteq\prod\limits_{n<\omega}\mu_n=\bigcup\limits_{i<\theta} X_i$, necessarily for some $i<\theta$, $\eta\in X_i$. Let $n\in(n(i),\omega)$ be such that $\beta_n$ is not in the closure of $\{\beta_\ell:\ell<\omega\mbox{ and }\ell\neq n\}$ and let $k>n$ be such that ${\bf t}^{n,k}_i=\mbox{ false}$. Now $\gamma=: F^{n,k}_i(\beta_0,\ldots,\beta_{n-1},\beta_{n+1},\ldots,\beta_{k-1})$ is well defined $\le\beta_n$ (as $\beta_n$ exemplifies that there is such $\beta$) and is $\ne \beta_n$ (by the choice of $\langle\beta_\ell:\ell<\omega\rangle$), so by the choice of $n(i)$ (so of $n$, $k$ and earlier of ${\bf t}^{n, k}_i$ and of $A_n$) we get contradiction to ``$\gamma<\beta_n$ are from $A_n$". \noindent 2)\ \ Let $M$ be an algebra with universe $\sum\limits_{n<\omega} \mu_n$ and $\le\theta$ functions (say $F^n_i$ for $i<\theta$, $n<\omega$, $F^n_i$ is $n$-place) exemplifying $\neg{\rm IND}'_\theta(\langle\mu_n:n<\omega \rangle)$. Let \[\Gamma=:\{\langle(k_n,i_n):n^*\le n<\omega\rangle:n^*<\omega\mbox{ and } \bigwedge_n n<k_n<\omega\mbox{ and }i_n<\theta\}.\] For $\rho=\langle(k_n,i_n):n^*\le n<\omega\rangle\in\Gamma$ let \[\begin{ALIGN} A_\rho=:&\big\{\eta\in\prod\limits_{n<\omega}\mu_n:\mbox{for every }n \in [n^*,\omega)\mbox{ we have}\\ &\qquad\eta(n)=F^{k_n-1}_{i_n}\left(\eta(0),\ldots,\eta(n-1),\eta(n+1), \ldots,\eta(k_n)\right)\big\}. \end{ALIGN}\] So, by the choice of $M$, $\prod\limits_{n<\omega}\mu_n=\bigcup\limits_{\rho \in\Gamma} A_\rho$. On the other hand, it is easy to check that $A_\rho\in J^1_{\bar \mu}$. \hfill$\square_{\ref{5.5}}$ \begin{Theorem} \label{5.6} If $\mu=\sum\limits_{n<\omega}\lambda_n$, $\lambda^{\aleph_0}_n<\lambda_{n+1}$ and $\mu<\lambda={\rm cf}(\lambda)<\mu^{+\omega}$\\ then ${\bf U}^{\aleph_0}_{I^0_{\langle\lambda_n:n<\omega\rangle}}(\lambda)= \lambda$ and even ${\bf U}^{\aleph_0}_{J^3_{\langle \lambda_n:n<\omega\rangle}} (\lambda)=\lambda$. \end{Theorem} \proof See \cite[\S6]{Sh:410}, \cite[\S2]{Sh:430}, and \cite{Sh:513} for considerably more. \begin{Lemma} \label{5.7} Assume $\lambda>2^{\aleph_0}$ and \begin{description} \item[$(*)$(a)] $\prod\limits_{n<\omega}\mu_n<\mu$ and $\mu^+<\lambda= {\rm cf}(\lambda)<\mu^{\aleph_0}$, \item[\ \ (b)] $\hat B_{\bar\mu}\notin I^0_{\bar\mu}$ and $\lim_n\sup\mu_n$ is infinite, \item[\ \ (c)] ${\bf U}^{\aleph_0}_{I^0_{\bar\mu}}(\lambda)=\lambda$ (note $I^0_{\bar \mu}$ is not required to be an ideal). \end{description} {\em Then} there is no universal member in ${\frak K}^{rs(p)}_\lambda$. \end{Lemma} \proof Let $S\subseteq\lambda$, $\bar C=\langle C_\delta:\delta\in S\rangle$ guesses clubs of $\lambda$, chosen as in the proof of \ref{3.3} (so $\alpha \in{\rm nacc}(C_\delta)\ \Rightarrow\ {\rm cf}(\alpha)>2^{\aleph_0}$). Instead of defining the relevant invariant we prove the theorem directly, but we could define it, somewhat cumbersomely (like \cite[III,\S3]{Sh:e}). Assume $H\in {\frak K}^{rs(p)}_\lambda$ is a pretender to universality; without loss of generality with the set of elements of $H$ equal to $\lambda$. Let $\chi=\beth_7(\lambda)^+$, ${\bar{\frak A}}=\langle {\frak A}_\alpha: \alpha<\lambda\rangle$ be an increasing continuous sequence of elementary submodels of $({\cal H}(\chi),\in,<^*_\chi)$, ${\bar {\frak A}}\restriction (\alpha+1)\in {\frak A}_{\alpha+1}$, $\|{\frak A}_\alpha\|<\lambda$, ${\frak A}_\alpha\cap\lambda$ an ordinal, ${\frak A}=\bigcup\limits_{\alpha<\lambda} {\frak A}_\alpha$ and $\{H,\langle\mu_n:n<\omega\rangle,\mu,\lambda\}\in {\frak A}_0$, so $B_{\bar \mu},\hat B_{\bar \mu} \in {\frak A}_0$ (where $\bar \mu=\langle\mu_n:n<\omega\rangle$, of course). For each $\delta\in S$, let ${\cal P}_\delta=:[C_\delta]^{\aleph_0}\cap {\frak A}$. Choose $A_\delta\subseteq C_\delta$ of order type $\omega$ almost disjoint from each $a\in {\cal P}_\delta$, and from $A_{\delta_1}$ for $\delta_1\in\delta\cap S$; its existence should be clear as $\lambda< \mu^{\aleph_0}$. So \begin{description} \item[$(*)_0$] every countable $A\in {\frak A}$ is almost disjoint to $A_\delta$. \end{description} By \ref{5.2}(2), $I^0_{\bar\mu}$ is $(<\aleph_1)$-based so by \ref{5.4}(1) and the assumption (c) we have \begin{description} \item[$(*)_1$] for every $f:\hat B_{\bar\mu}\longrightarrow\lambda$ for some countable $Y\subseteq \hat B_{\bar\mu}$, $Y\notin I^0_{\bar\mu}$, we have $f\restriction Y\in {\frak A}$ \end{description} (remember $(\prod\limits_{n<\omega}\mu_n)^{\aleph_0}=\prod\limits_{n<\omega} \mu_n$). \noindent Let $B$ be $\bigoplus\{G^n_{\alpha,i}:n<\omega,\alpha<\lambda,\,i< \sum\limits_{k<\omega}\mu_k\}$, where \[G^n_{\alpha,i}=\langle x^n_{\alpha,i}\rangle_{G^n_{\alpha,i}}\cong{\Bbb Z}/p^{n+1}{\Bbb Z}.\] \noindent So $B$, $\hat B$, $\langle(n,\alpha,i,x^n_{\alpha,i}):n<\omega,\alpha<\lambda, i<\sum\limits_{k<\omega}\mu_k\rangle$ are well defined. Let $G$ be the subgroup of $\hat B$ generated by: \[\begin{ALIGN} B\cup\big\{x\in\hat B: &\mbox{for some }\delta\in S,\, x\mbox{ is in the closure of }\\ &\bigoplus\{G^n_{\alpha,i}:n<\omega,i<\mu_n,\alpha\mbox{ is the }n\mbox{th element of } A_\delta\}\big\}. \end{ALIGN}\] As $\prod\limits_{n<\omega}\mu_n<\mu<\lambda$, clearly $G\in {\frak K}^{rs(p)}_\lambda$, without loss of generality the set of elements of $G$ is $\lambda$ and let $h:G\longrightarrow H$ be an embedding. Let \[E_0=:\{\delta<\lambda:({\frak A}_\delta,h \restriction \delta,\;G \restriction\delta)\prec({\frak A},h,G)\},\] \[E=:\{\delta<\lambda:{\rm otp}(E_0\cap\delta)=\delta\}.\] They are clubs of $\lambda$, so for some $\delta\in S$, $C_\delta\subseteq E$ (and $\delta\in E$ for simplicity). Let $\eta_\delta$ enumerate $A_\delta$ increasingly. There is a natural embedding $g = g_\delta$ of $B_{\bar \mu}$ into $G$: \[g({}^* t^n_i) = x^n_{\eta_\delta(n),i}.\] Let $\hat g_\delta$ be the unique extension of $g_\delta$ to an embedding of $\hat B_{\bar\mu}$ into $G$; those embeddings are pure, (in fact $g''_\delta (\hat B_{\bar\mu})\setminus g''_\delta(B_\mu)\subseteq G\setminus G\cap {\frak A}_\delta$). So $h\circ\hat g_\delta$ is an embedding of $\hat B_{\bar \mu}$ into $H$, not necessarily pure but still an embedding, so the distance function can become smaller but not zero and \[h\circ\hat g_\delta(\hat B_{\bar\mu})\setminus h\circ g_\delta(B_\mu) \subseteq H\setminus {\frak A}_\delta.\] Remember $\hat B_{\bar\mu}\subseteq {\frak A}_0$ (as it belongs to ${\frak A}_0$ and has cardinality $\prod\limits_{n<\omega}\mu_n<\lambda$ and $\lambda\cap {\frak A}_0$ is an ordinal). By $(*)_1$ applied to $f=h\circ\hat g$ there is a countable $Y \subseteq \hat B_{\bar \mu}$ such that $Y \notin I^0_{\bar\mu}$ and $f \restriction Y \in {\frak A}$. But, from $f \restriction Y$ we shall below reconstruct some countable sets not almost disjoint to $A_\delta$, reconstruct meaning in ${\frak A}$, in contradiction to $(*)_0$ above. As $Y\notin I^0_{\bar \mu}$ we can find an infinite $S^*\subseteq\omega\setminus m^*$ and for $n\in S^*$, $z_n\in\bigoplus\limits_{\alpha<\mu_n} K^n_\alpha\setminus\{0\}$ and $y^\ell_n\in\hat B_{\bar\mu}$ (for $\ell<\omega$) such that: \begin{description} \item[$(*)_2$] $z_n+y_{n,\ell}\in\langle Y\rangle_{\hat B_{\bar\mu}}$,\qquad and \item[$(*)_3$] $y_{n,\ell}\in p^\ell\,\hat B_{\bar\mu}$. \end{description} Without loss of generality $pz_n=0\ne z_n$ hence $p\,y^\ell_n=0$. Let \[\nu_\delta(n)=:\min(C_\delta\setminus (\eta_\delta(n)+1)),\quad z^*_n= (h\circ\hat g_\delta)(z_n)\quad\mbox{ and }\quad y^*_{n,\ell}=(h\circ\hat g_\delta)(y_{n,\ell}).\] Now clearly $\hat g_\delta(z_n)=g_\delta(z_n)=x^n_{\eta_\delta(n),i}\in G\restriction\nu_\delta(n)$, hence $(h\circ\hat g_\delta)(z_n)\notin H \restriction\eta_\delta(n)$, that is $z^*_n\notin H\restriction\eta_\delta(n)$. So $z^*_n\in H_{\nu_\delta(n)}\setminus H_{\eta_\delta(n)}$ belongs to the $p$-adic closure of ${\rm Rang}(f\restriction Y)$. As $H$, $G$, $h$ and $f\restriction Y$ belongs to ${\frak A}$, also $K$, the closure of ${\rm Rang}(f\restriction Y)$ in $H$ by the $p$-adic topology belongs to ${\frak A}$, and clearly $|K|\leq 2^{\aleph_0}$, hence \[A^*=\{\alpha\in C_\delta: K\cap H_{\min(C_\delta\setminus (\alpha+1))}\setminus H_\alpha \mbox{ is not empty}\}\] is a subset of $C_\delta$ of cardinality $\leq 2^{\aleph_0}$ which belongs to ${\frak A}$, hence $[A^*]^{\aleph_0}\subseteq {\frak A}$ but $A_\delta\subseteq A^*$ so $A_\delta\in {\frak A}$, a contradiction. \hfill$\square_{\ref{5.7}}$ \section{Implications between the existence of universals} \begin{Theorem} \label{6.1} Let $\bar n=\langle n_i:i<\omega\rangle$, $n_i\in [1,\omega)$. Remember \[J^2_{\bar n}=\{A\subseteq\prod_{i<\omega} n_i:A \mbox{ is nowhere dense}\}.\] Assume $\lambda\ge 2^{\aleph_0}$, $T^{\aleph_0}_{J^3_{\bar n}}(\lambda)= \lambda$ or just $T^{\aleph_0}_{J^2_{\bar n}}(\lambda)=\lambda$ for every such $\bar n$, and \[n<\omega\quad\Rightarrow\quad\lambda_n\le\lambda_{n+1}\le\lambda_\omega =\lambda\quad\mbox{ and}\] \[\lambda\le\prod_{n<\omega}\lambda_n\quad\mbox{ and }\quad\bar\lambda= \langle\lambda_i:i\le\omega\rangle.\] \begin{enumerate} \item If in ${\frak K}^{fc}_{\bar \lambda}$ there is a universal member {\em then} in ${\frak K}^{rs(p)}_\lambda$ there is a universal member. \item If in ${\frak K}^{fc}_\lambda$ there is a universal member for ${\frak K}^{fc}_{\bar \lambda}$ {\em then} in \[{\frak K}^{rs(p)}_{\bar\lambda}=:\{G\in {\frak K}^{rs(p)}_\lambda:\lambda_n (G)\le\lambda\}\] there is a universal member (even for ${\frak K}^{rs(p)}_\lambda$). \end{enumerate} ($\lambda_n(G)$ were defined in \ref{1.1}). \end{Theorem} \begin{Remark} \begin{enumerate} \item Similarly for ``there are $M_i\in {\frak K}_{\lambda_1}$ ($i<\theta$) with $\langle M_i:i<\theta\rangle$ being universal for ${\frak K}_\lambda$''. \item The parallel of \ref{1.1} holds for ${\frak K}^{fc}_\lambda$. \item By \S5 only the case $\lambda$ singular or $\lambda=\mu^+\ \&\ {\rm cf}(\mu)=\aleph_0\ \& \ (\forall \alpha< \mu)(|\alpha|^{\aleph_0}<\mu)$ is of interest for \ref{6.1}. \end{enumerate} \end{Remark} \proof 1)\ \ By \ref{1.1}, (2) $\Rightarrow$ (1). More elaborately, by part (2) of \ref{6.1} below there is $H\in {\frak K}^{rs(p)}_{\bar \lambda}$ which is universal in ${\frak K}^{rs(p)}_{\bar \lambda}$. Clearly $|G|=\lambda$ so $H\in {\frak K}^{rs(p)}_\lambda$, hence for proving part (1) of \ref{6.1} it suffices to prove that $H$ is a universal member of ${\frak K}^{rs(p)}_\lambda$. So let $G\in {\frak K}^{rs(p)}_\lambda$, and we shall prove that it is embeddable into $H$. By \ref{1.1} there is $G'$ such that $G\subseteq G'\in {\frak K}^{rs(p)}_{\bar \lambda}$. By the choice of $H$ there is an embedding $h$ of $G'$ into $H$. So $h\restriction G$ is an embedding of $G$ into $H$, as required. \noindent 2)\ \ Let $T^*$ be a universal member of ${\frak K}^{fc}_{\bar \lambda}$ (see \S2) and let $P_\alpha = P^{T^*}_\alpha$. Let $\chi>2^\lambda$. Without loss of generality $P_n=\{n\}\times \lambda_n$, $P_\omega=\lambda$. Let \[B_0=\bigoplus\{G^n_t:n<\omega,t\in P_n \},\] \[B_1=\bigoplus \{G^n_t: n< \omega\mbox{ and }t\in P_n\},\] where $G^n_t\cong {\Bbb Z}/p^{n+1}{\Bbb Z}$, $G^n_t$ is generated by $x^n_t$. Let ${\frak B}\prec ({\cal H}(\chi),\in,<^*_\chi)$, $\|{\frak B}\|= \lambda$, $\lambda+1\subseteq {\frak B}$, $T^*\in {\frak B}$, hence $B_0$, $B_1\in {\frak B}$ and $\hat B_0, \hat B_1\in {\frak B}$ (the torsion completion of $B$). Let $G^* =\hat B_1\cap {\frak B}$. Let us prove that $G^*$ is universal for ${\frak K}^{rs(p)}_{\bar \lambda}$ (by \ref{1.1} this suffices). Let $G \in {\frak K}^{rs(p)}_{\lambda}$, so by \ref{1.1} without loss of generality $B_0 \subseteq G\subseteq\hat B_0$. We define $R$: \[\begin{ALIGN} R=\big\{\eta:&\eta\in\prod\limits_{n<\omega}\lambda_n\mbox{ and for some } x\in G\mbox{ letting }\\ &x=\sum\{a^n_i\,p^{n-k}\,x^n_i:n<\omega,i\in w_n(x)\}\mbox{ where }\\ &w_n(x)\in [\lambda_n]^{<\aleph_0},a^n_i\,p^{n-k}\,x^n_i\ne 0\mbox{ we have }\\ &\bigwedge\limits_n\eta(n)\in w_n(x)\cup \{\ell: \ell+|w_n(x)|\leq n\}\big\}. \end{ALIGN}\] Lastly let $M =:(R\cup\bigcup\limits_{n<\omega}\{n\}\times\lambda_n,\,P_n,\, F_n)_{n<\omega}$ where $P_n=\{n\}\times\lambda_n$ and $F_n(\eta)=(n,\eta(n))$, so clearly $M\in {\frak K}^{fc}_{\bar\lambda}$. Consequently, there is an embedding $g:M\longrightarrow T^*$, so $g$ maps $\{n\}\times\lambda_n$ into $P^{T^*}_n$ and $g$ maps $R$ into $P^{T^*}_\omega$. Let $g(n,\alpha)= (n,g_n(\alpha))$ (i.e. this defines $g_n$). Clearly $g\restriction (\cup P^M_n)=g\restriction (\bigcup\limits_n\{n\}\times\lambda_n)$ induces an embedding $g^*$ of $B_0$ to $B_1$ (by mapping the generators into the generators). \noindent The problem is why: \begin{description} \item[$(*)$] if $x=\sum\{a^n_i\,p^{n-k}\,x^n_i:n<\omega,i\in w_n(x)\}\in G$ {\em then} $g^*(x)=\sum\{a^n_i\,p^{n-k}\,g^*(x^n_i):n<\omega,i\in w_n(x)\}\in G^*$. \end{description} As $G^*=\hat B_1\cap {\frak B}$, and $2^{\aleph_0}+1\subseteq {\frak B}$, it is enough to prove $\langle g^{\prime\prime}(w_n(x)):n<\omega\rangle\in {\frak B}$. Now for notational simplicity $\bigwedge\limits_n [|w_n(x)|\ge n+1]$ (we can add an element of $G^*\cap {\frak B}$ or just repeat the arguments). For each $\eta \in\prod\limits_{n<\omega} w_n(x)$ we know that \[g(\eta)=\langle g(\eta(n)):n<\omega\rangle\in T^*\quad\mbox{ hence is in } {\frak B}\] (as $T^*\in {\frak B}$, $|T^*|\le\lambda$). Now by assumption there is $A\subseteq\prod\limits_{n<\omega} w_n(x)$ which is not nowhere dense such that $g \restriction A\in {\frak B}$, hence for some $n^*$ and $\eta^* \in \prod\limits_{\ell<n^*}w_\ell(x)$, $A$ is dense above $\eta^*$ (in $\prod\limits_{n<\omega} w_n(x)$). Hence \[\langle\{\eta(n):\eta\in A\}:n^* \le n<\omega\rangle=\langle w_n[x]:n^*\le n<\omega\rangle,\] but the former is in ${\frak B}$ as $A\in {\frak B}$, and from the latter the desired conclusion follows. \hfill$\square_{\ref{6.1}}$ \section{Non-existence of universals for trees with small density} For simplicity we deal below with the case $\delta=\omega$, but the proof works in general (as for ${\frak K}^{fr}_{\bar\lambda}$ in \S2). Section 1 hinted we should look at ${\frak K}^{tr}_{\bar\lambda}$ not only for the case $\bar\lambda=\langle\lambda:\alpha\le\omega\rangle$ (i.e. ${\frak K}^{tr}_\lambda$), but in particular for \[\bar\lambda=\langle\lambda_n:n<\omega\rangle\char 94\langle\lambda\rangle, \qquad \lambda^{\aleph_0}_n<\lambda_{n+1}<\mu<\lambda={\rm cf}(\lambda)< \mu^{\aleph_0}.\] Here we get for this class (embeddings are required to preserve levels), results stronger than the ones we got for the classes of Abelian groups we have considered. \begin{Theorem} \label{7.1} Assume that \begin{description} \item[(a)] $\bar\lambda=\langle\lambda_\alpha:\alpha\le\omega\rangle$, $\lambda_n<\lambda_{n+1}<\lambda_\omega$, $\lambda=\lambda_\omega$, all are regulars, \item[(b)] $D$ is a filter on $\omega$ containing cobounded sets, \item[(c)] ${\rm tcf}(\prod \lambda_n /D)=\lambda$ (indeed, we mean $=$, we could just use $\lambda\in{\rm pcf}_D(\{\lambda_n:n<\omega\})$), \item[(d)] $(\sum\limits_{n<\omega}\lambda_n)^+<\lambda<\prod\limits_{n< \omega}\lambda_n$. \end{description} {\em Then} there is no universal member in ${\frak K}^{tr}_{\bar\lambda}$. \end{Theorem} \proof We first notice that there is a sequence $\bar P=\langle P_\alpha: \sum\limits_{n<\omega}\lambda_n<\alpha<\lambda\rangle$ such that: \begin{enumerate} \item $|P_\alpha|<\lambda$, \item $a\in P_\alpha\quad\Rightarrow\quad a$ is a closed subset of $\alpha$ of order type $\leq\sum\limits_{n<\omega}\lambda_n$, \item $a\in\bigcup\limits_{\alpha<\lambda} P_\alpha\ \&\ \beta\in{\rm nacc}(a) \quad \Rightarrow\quad a\cap\beta\in P_\beta$, \item For all club subsets $E$ of $\lambda$, there are stationarily many $\delta$ for which there is an $a\in\bigcup\limits_{\alpha<\lambda} P_\alpha$ such that \[{\rm cf}(\delta)=\aleph_0\ \&\ a\in P_\delta\ \&\ {\rm otp}(a)=\sum_{n<\omega} \lambda_n\ \&\ a\subseteq E.\] \end{enumerate} [Why? If $\lambda=(\sum\limits_{n<\omega}\lambda_n)^{++}$, then it is the successor of a regular, so we use \cite[\S4]{Sh:351}, i.e. \[\{\alpha<\lambda:{\rm cf}(\alpha)\le(\sum_{n<\omega}\lambda_n)\}\] is the union of $(\sum\limits_{n<\omega}\lambda_n)^+$ sets with squares.\\ If $\lambda>(\sum\limits_{n<\omega}\lambda_n)^{++}$, then we can use \cite[\S1]{Sh:420}, which guarantees that there is a stationary $S\in I[\lambda]$.] We can now find a sequence \[\langle f_\alpha,g_{\alpha,a}:\alpha<\lambda,a\in P_\alpha\rangle\] such that: \begin{description} \item[(a)] $\bar f=\langle f_\alpha:\alpha<\lambda\rangle$ is a $<_D$-increasing cofinal sequence in $\prod\limits_{n<\omega}\lambda_n$, \item[(b)] $g_{\alpha,a}\in\prod\limits_{n<\omega}\lambda_n$, \item[(c)] $\bigwedge\limits_{\beta<\alpha} f_\beta<_D g_{\alpha,a}<_D f_{\alpha+1}$, \item[(d)] $\lambda_n>|a|\ \&\ \beta\in{\rm nacc}(a)\quad \Rightarrow\quad g_{\beta,a\cap\beta}(n)<g_{\alpha,a}(n)$. \end{description} [How? Choose $\bar f$ by ${\rm tcf}(\prod\limits_{n<\omega}\lambda_n/D)=\lambda$. Then choose $g$'s by induction, possibly throwing out some of the $f$'s; this is from \cite[II, \S1]{Sh:g}.] Let $T\in {\frak K}^{tr}_{\bar \lambda}$. We introduce for $x\in{\rm lev}_\omega(T)$ and $\ell<\omega$ the notation $F^T_\ell(x) = F_\ell(x)$ to denote the unique member of ${\rm lev}_\ell(T)$ which is below $x$ in the tree order of $T$. \noindent For $a\in\bigcup\limits_{\alpha<\lambda} P_\alpha$, let $a=\{ \alpha_{a,\xi}:\xi<{\rm otp}(a)\}$ be an increasing enumeration. We shall consider two cases. In the first one, we assume that the following statement $(*)$ holds. In this case, the proof is easier, and maybe $(*)$ always holds for some $D$, but we do not know this at present. \begin{description} \item[{{$(*)$}}] There is a partition $\langle A_n:n < \omega \rangle$ of $\omega$ into sets not disjoint to any member of $D$. \end{description} In this case, let for $n\in\omega$, $D_n$ be the filter generated by $D$ and $A_n$. Let for $a\in\bigcup\limits_{\alpha<\lambda} P_\alpha$ with ${\rm otp}(a)= \sum\limits_{n<\omega}\lambda_n$, and for $x\in{\rm lev}_\omega(T)$, \[{\rm inv}(x,a,T)=:\langle\xi_n(x,a,T):n<\omega\rangle,\] where \[\begin{ALIGN} \xi_n(x,a,T)=:\min\big\{\xi<{\rm otp}(a)\!:&\mbox{ for some }m<\omega \mbox{ we have }\\ & \langle F^T_\ell(x): \ell<\omega\rangle<_{D_n} g_{\alpha', a'}\mbox{ where }\\ &\alpha'=\alpha_{a,\omega\xi+m}\mbox{ and } a'= a\cap\alpha'\big\}. \end{ALIGN}\] Let \[{\rm INv}(a,T)=:\{{\rm inv}(x,a,T):x\in T\ \&\ {\rm lev}_T(x)=\omega\},\] \[\begin{ALIGN} {\rm INV}(T)=:\big\{c:&\mbox{for every club } E\subseteq\lambda,\mbox{ for some } \delta\mbox{ and } a\in P\\ &\mbox{we have }{\rm otp}(a)=\sum\lambda_n\ \&\ a\subseteq E\ \&\ a\in P_\delta\\ &\mbox{and for some } x\in T\mbox{ of } {\rm lev}_T(x)=\omega,\ c={\rm inv}(x,a,T) \big\}. \end{ALIGN}\] (Alternatively, we could have looked at the function giving each $a$ the value ${\rm INv}(a,T)$, and then divide by a suitable club guessing ideal as in the proof in \S3, see Definition \ref{3.7}.) \noindent Clearly \medskip \noindent{\bf Fact}:\hspace{0.15in} ${\rm INV}(T)$ has cardinality $\le\lambda$. \medskip The main point is the following \medskip \noindent{\bf Main Fact}:\hspace{0.15in} If ${\bf h}:T^1\longrightarrow T^2$ is an embedding, {\em then\/} \[{\rm INV}(T^1)\subseteq{\rm INV}(T^2).\] \medskip \noindent{\em Proof of the {\bf Main Fact} under $(*)$}\ \ \ We define for $n \in\omega$ \[E_n=:\big\{\delta<\lambda_n:\,\delta>\bigcup_{\ell<n}\lambda_\ell\mbox{ and }\left(\forall x\in{\rm lev}_n(T^1)\right)({\bf h}(x)<\delta \Leftrightarrow x<\delta) \big\}.\] We similarly define $E_\omega$, so $E_n$ ($n\in\omega$) and $E_\omega$ are clubs (of $\lambda_n$ and $\lambda$ respectively). Now suppose $c\in{\rm INV}(T_1) \setminus{\rm INV}(T_2)$. Without loss of generality $E_\omega$ is (also) a club of $\lambda$ which exemplifies that $c\notin{\rm INV}(T_2)$. For $h\in \prod\limits_{n<\omega}\lambda_n$, let \[h^+(n)=:\min(E_n\setminus h(n)),\quad\mbox{ and }\quad\beta[h]=\min\{\beta <\lambda:h<f_\beta\}.\] (Note that $h<f_{\beta[h]}$, not just $h<_D f_{\beta[h]}$.) For a sequence $\langle h_i:i<i^*\rangle$ of functions from $\prod\limits_{n<\omega} \lambda_n$, we use $\langle h_i:i<i^* \rangle^+$ for $\langle h^+_i:i<i^* \rangle$. Now let \[E^*=:\big\{\delta<\lambda:\mbox{if }\alpha<\delta\mbox{ then } \beta[f^+_\alpha]<\delta\mbox{ and }\delta\in{\rm acc}(E_\omega)\big\}.\] Thus $E^*$ is a club of $\lambda$. Since $c\in{\rm INV}(T_1)$, there is $\delta< \lambda$ and $a\in P_\delta$ such that for some $x\in{\rm lev}_\omega(T_1)$ we have \[a\subseteq E^*\ \&\ {\rm otp}(a)=\sum_{n<\omega}\lambda_n\ \&\ c={\rm inv}(x,a,T_1).\] Let for $n\in\omega$, $\xi_n=:\xi_n(x,a,T_1)$, so $c=\langle\xi_n:n<\omega \rangle$. Also let for $\xi<\sum\limits_{n<\omega}\lambda_n$, $\alpha_\xi=: \alpha_{a,\xi}$, so $a=\langle\alpha_\xi:\xi<\sum\limits_{n<\omega}\lambda_n \rangle$ is an increasing enumeration. Now fix an $n<\omega$ and consider ${\bf h}(x)$. Then we know that for some $m$ \begin{description} \item[($\alpha$)] $\langle F^{T_1}_\ell(x):\ell<\omega\rangle<_{D_n} g_{\alpha'}$ where $\alpha'=\alpha_{\omega\xi_n+m}$\qquad and \item[($\beta$)] for no $\xi<\xi_n$ is there such an $m$. \end{description} Now let us look at $F^{T_1}_\ell(x)$ and $F^{T_2}_\ell({\bold h}(x))$. They are not necessarily equal, but \begin{description} \item[($\gamma$)] $\min(E_\ell\setminus F^{T_1}_\ell(x))=\min(E_\ell \setminus F^{T_2}_\ell({\bf h}(x))$ \end{description} (by the definition of $E_\ell$). Hence \begin{description} \item[($\delta$)] $\langle F^{T_1}_\ell(x):\ell<\omega\rangle^+=\langle F^{T_2}_\ell({\bf h}(x)):\ell<\omega\rangle^+$. \end{description} Now note that by the choice of $g$'s \begin{description} \item[($\varepsilon$)] $(g_{\alpha_\varepsilon,a\cap\alpha_\varepsilon})^+ <_{D_n} g_{\alpha_{\varepsilon+1},a\cap\alpha_{\varepsilon+1}}$. \end{description} \relax From $(\delta)$ and $(\varepsilon)$ it follows that $\xi_n({\bf h}(x),a, T^2)=\xi_n(x,a,T^1)$. Hence $c\in{\rm INV}(T^2)$. \hfill$\square_{\mbox{Main Fact}}$ \medskip Now it clearly suffices to prove: \medskip \noindent{\bf Fact A:}\hspace{0.15in} For each $c=\langle\xi_n:n<\omega \rangle\in {}^\omega(\sum\limits_{n<\omega}\lambda_n)$ we can find a $T\in {\frak K}^{tr}_{\bar\lambda}$ such that $c\in{\rm INV}(T)$. \medskip \noindent{\em Proof of the Fact A in case $(*)$ holds}\ \ \ For each $a\in \bigcup\limits_{\delta<\lambda} P_\delta$ with ${\rm otp}(a)=\sum\limits_{n\in \omega}\lambda_n$ we define $x_{c,a}=:\langle x_{c,a}(\ell):\ell<\omega \rangle$ by: \begin{quotation} if $\ell\in A_n$, then $x_{c,a}(\ell)=\alpha_{a,\omega\xi_n+\delta}$. \end{quotation} Let \[T=\bigcup_{n<\omega}\prod_{\ell<n}\lambda_\ell\cup\big\{x_{c,a}:a\in \bigcup_{\delta<\lambda} P_\delta\ \&\ {\rm otp}(a)=\sum_{n<\omega}\lambda_n \big\}.\] We order $T$ by $\triangleleft$. It is easy to check that $T$ is as required. \hfill$\square_A$ \medskip Now we are left to deal with the case that $(*)$ does not hold. Let \[{\rm pcf}(\{\lambda_n:n<\omega\})=\{\kappa_\alpha:\alpha\le\alpha^*\}\] be an enumeration in increasing order so in particular \[\kappa_{\alpha^*}=\max{\rm pcf}(\{\lambda_n:n<\omega\}).\] Without loss of generality $\kappa_{\alpha^*}=\lambda$ (by throwing out some elements if necessary) and $\lambda\cap{\rm pcf}(\{\lambda_n:n<\omega\})$ has no last element (this appears explicitly in \cite{Sh:g}, but is also straightforward from the pcf theorem). In particular, $\alpha^*$ is a limit ordinal. Hence, without loss of generality \[D=\big\{A\subseteq\omega:\lambda>\max{\rm pcf}\{\lambda_n:n\in\omega\setminus A\} \big\}.\] Let $\langle {\frak a}_{\kappa_\alpha}:\alpha\le\alpha^*\rangle$ be a generating sequence for ${\rm pcf}(\{\lambda_n:n<\omega\})$, i.e. \[\max{\rm pcf}({\frak a}_{\kappa_\alpha})=\kappa_\alpha\quad\mbox{ and }\quad \kappa_\alpha\notin{\rm pcf}(\{\lambda_n:n<\omega\}\setminus {\frak a}_{\kappa_\alpha}).\] (The existence of such a sequence follows from the pcf theorem). Without loss of generality, \[{\frak a}_{\alpha^*} = \{ \lambda_n:n < \omega\}.\] Now note \begin{Remark} If ${\rm cf}(\alpha^*)=\aleph_0$, then $(*)$ holds. \end{Remark} Why? Let $\langle\alpha(n):n<\omega\rangle$ be a strictly increasing cofinal sequence in $\alpha^*$. Let $\langle B_n:n<\omega\rangle$ partition $\omega$ into infinite pairwise disjoint sets and let \[A_\ell=:\big\{k<\omega:\bigvee_{n\in B_\ell}[\lambda_k\in {\frak a}_{\kappa_{\alpha(n)}}\setminus\bigcup_{m<n} {\frak a}_{\kappa_{\alpha(m)}}] \big\}.\] To check that this choice of $\langle A_\ell:\ell<\omega\rangle$ works, recall that for all $\alpha$ we know that $\alpha_{\kappa_\alpha}$ does not belong to the ideal generated by $\{{\frak a}_{\kappa_\beta}: \beta<\alpha\}$ and use the pcf calculus. \hfill$\square$ Now let us go back to the general case, assuming ${\rm cf}(\alpha^*)>\aleph_0$. Our problem is the possibility that \[{\cal P}(\{\lambda_n:n<\omega\})/J_{<\lambda}[\{\lambda_n:n<\omega\}].\] is finite. Let now $A_\alpha=:\{n:\lambda_n\in {\frak a}_\alpha\}$, and \[\begin{array}{lll} J_\alpha &=: &\big\{A\subseteq\omega:\max{\rm pcf}\{\lambda_\ell:\ell\in A\}< \kappa_\alpha\big\}\\ J'_\alpha &=: &\big\{A\subseteq\omega:\max{\rm pcf}\{\lambda_\ell:\ell\in A\}\cap {\frak a}_{\kappa_\alpha}<\kappa_\alpha\big\}. \end{array}\] We define for $T\in {\frak K}^{tr}_{\bar\lambda}$, $x\in{\rm lev}_\omega(T)$, $\alpha<\alpha^*$ and $a\in\bigcup\limits_{\delta<\lambda} P_\delta$: \[\begin{ALIGN} \xi^*_\alpha(x,a,T)=:\min\big\{\xi:&\bigvee_m [\langle F^T_\ell(x):\ell< \omega\rangle <_{J'_\alpha} g_{\alpha', a'}\mbox{ where }\\ &\qquad \alpha'=\alpha_{a,\omega \xi+m}\mbox{ and }a'=a\cap \alpha'\big\}. \end{ALIGN} \] Let \[{\rm inv}_\alpha(x,a,T)=:\langle\xi^*_{\alpha+n}(x,a,T):n<\omega\rangle,\] \[{\rm INv}(a,T)=:\big\{{\rm inv}_\alpha(x,a,T):x\in T\ \&\ \alpha<\alpha^* \ \&\ {\rm lev}_T(x)=\omega\big\},\] and \[\begin{ALIGN} {\rm INV}(T)=\big\{c:&\mbox{for every club } E^*\mbox{ of }\lambda\mbox{ for some } a\in\bigcup\limits_{\delta<\lambda} P_\delta\\ &\mbox{with }{\rm otp}(a)=\sum\limits_{n<\omega}\lambda_n\mbox{ for arbitrarily large }\alpha<\alpha^*,\\ &\mbox{there is }x\in{\rm lev}_\omega(T)\mbox{ such that }{\rm inv}_\alpha(x,a,T)= c\big\}. \end{ALIGN}\] As before, the point is to prove the Main Fact. \medskip \noindent{\em Proof of the {\bf Main Fact} in general}\ \ \ Suppose ${\bf h}: T^1\longrightarrow T^2$ and $c\in{\rm INV}(T^1)\setminus{\rm INV}(T^2)$. Let $E'$ be a club of $\lambda$ which witnesses that $c\notin{\rm INV}(T^2)$. We define $E_n,E_\omega$ as before, as well as $E^*$ ($\subseteq E_\omega\cap E'$). Now let us choose $a\in \bigcup\limits_{\delta<\lambda} P_\delta$ with $a\subseteq E^*$ and ${\rm otp}(a)= \sum\limits_{n<\omega}\lambda_n$. So $a=\{\alpha_{a,\xi}:\xi<\sum\limits_{n< \omega}\lambda_n\}$, which we shorten as $a=\{\alpha_\xi:\xi<\sum\limits_{n< \omega}\lambda_n\}$. For each $\xi<\sum\limits_{n<\omega}\lambda_n$, as before, we know that \[(g_{\alpha_\xi,a\cap\alpha_\xi})^+<_{J^*_\alpha}g_{\alpha_{\xi+1},a\cap \alpha_{\xi+1}}.\] Therefore, there are $\beta_{\xi,\ell}<\alpha^*$ ($\ell<\ell_\xi$) such that \[\{\ell:g^+_{\alpha_\xi,a\cap\alpha_\xi}(\ell)\ge g_{\alpha_{\xi+1},a\cap \alpha_{\xi+1}}(\ell)\}\supseteq\bigcup_{\ell<\ell_\xi} A_{\beta_{\xi,\ell}}.\] Let $c=\langle\xi_n:n<\omega\rangle$ and let \[\Upsilon=\big\{\beta_{\xi,\ell}:\mbox{for some } n\mbox{ and } m\mbox{ we have }\xi=\omega\xi_n+m\ \&\ \ell<\omega\big\}.\] Thus $\Upsilon\subseteq\alpha^*$ is countable. Since ${\rm cf}(\alpha^*)>\aleph_0$, the set $\Upsilon$ is bounded in $\alpha^*$. Now we know that $c$ appears as an invariant for $a$ and arbitrarily large $\delta<\alpha^*$, for some $x_{a,\delta}\in{\rm lev}_\omega(T_1)$. If $\delta>\sup(\Upsilon)$, $c\in{\rm INV}(T^2)$ is exemplified by $a,\delta,{\bf h}(x_{\alpha,\delta})$, just as before. \hfill$\square$ \medskip We still have to prove that every $c=\langle\xi_n:n<\omega\rangle$ appears as an invariant; i.e. the parallel of Fact A. \medskip \noindent{\em Proof of Fact A in the general case:}\ \ \ Define for each $a\in \bigcup\limits_{\delta<\lambda} P_\delta$ with ${\rm otp}(a)=\sum\limits_{n<\omega} \lambda_n$ and $\beta<\alpha^*$ \[x_{c,a,\beta}=\langle x_{c,a,\beta}(\ell):\ell<\omega\rangle,\] where \[x_{c,a,\beta}(\ell)=\left\{ \begin{array}{lll} \alpha_{a,\omega\xi_n+\delta} &\mbox{\underbar{if}} &\lambda_\ell\in {\frak a}_{\beta+k}\setminus\bigcup\limits_{k'<k} {\frak a}_{\beta+k'}\\ 0 &\mbox{\underbar{if}} &\lambda_\ell\notin {\frak a}_{\beta+k}\mbox{ for any } k<\omega. \end{array}\right.\] Form the tree as before. Now for any club $E$ of $\lambda$, we can find $a\in \bigcup\limits_{\delta<\lambda} P_\delta$ with ${\rm otp}(a)=\sum\limits_{n<\omega} \lambda_n$, $a\subseteq E$ such that $\langle x_{c,a,\beta}:\beta<\alpha^* \rangle$ shows that $c\in{\rm INV}(T)$. \hfill$\square_{\ref{7.1}}$ \begin{Remark} \begin{enumerate} \item Clearly, this proof shows not only that there is no one $T$ which is universal for ${\frak K}^{tr}_{\bar\lambda}$, but that any sequence of $<\prod\limits_{n<\omega}\lambda_n$ trees will fail. This occurs generally in this paper, as we have tried to mention in each particular case. \item The case ``$\lambda<2^{\aleph_0}$" is included in the theorem, though for the Abelian group application the $\bigwedge\limits_{n<\omega} \lambda^{\aleph_0}_n<\lambda_{n+1}$ is necessary. \end{enumerate} \end{Remark} \begin{Remark} \label{7.1A} \begin{enumerate} \item If $\mu^+< \lambda={\rm cf}(\lambda)<\chi<\mu^{\aleph_0}$ and $\chi^{[\lambda]}<\mu^{\aleph_0}$ (or at least $T_{{\rm id}^a(\bar C)}(\chi)<\mu^{\aleph_0}$) we can get the results for ``no $M \in {\frak K}^x_\chi$ is universal for ${\frak K}^x_\lambda$", see \S8 (and \cite{Sh:456}). \end{enumerate} \end{Remark} \noindent We can below (and subsequently in \S8) use $J^3_{\bar m}$ as in \S6. \begin{Theorem} \label{7.2} Assume that $2^{\aleph_0}<\lambda_0$, $\bar\lambda=\langle\lambda_n:n<\omega \rangle\char 94\langle\lambda\rangle$, $\mu=\sum\limits_{n<\omega}\lambda_n$, $\lambda_n<\lambda_{n+1}$, $\mu^+<\lambda={\rm cf}(\lambda)<\mu^{\aleph_0}$.\\ \underbar{If}, for simplicity, $\bar m=\langle m_i:i<\omega\rangle=\langle \omega:i<\omega\rangle$ (actually $m_i\in [2,\omega]$ or even $m_i\in [2,\lambda_0)$, $\lambda_0<\lambda$ are O.K.) and ${\bf U}^{<\mu}_{J^2_{\bar m}} (\lambda)=\lambda$ (remember \[J^2_{\bar m}=\{A\subseteq\prod_{i<\omega} m_i:A \mbox{ is nowhere dense}\}\] and definition \ref{5.3}),\\ \underbar{then} in ${\frak K}^{tr}_{\bar\lambda}$ there is no universal member. \end{Theorem} \proof 1)\ \ Let $S\subseteq\lambda$, $\bar C=\langle C_\delta:\delta\in S \rangle$ be a club guessing sequence on $\lambda$ with ${\rm otp}(C_\delta)\ge\sup \lambda_n$. We assume that we have ${\bar{\frak A}}=\langle {\frak A}_\alpha: \alpha<\lambda\rangle$, $J^2_{\bar m}$, $T^*\in {\frak A}_0$ ($T^*$ is a candidate for the universal), $\bar C=\langle C_\delta:\delta\in S\rangle\in {\frak A}_\alpha$, ${\frak A}_\alpha\prec({\cal H}(\chi),\in,<^*_\chi)$, $\chi= \beth_7(\lambda)^+$, $\|{\frak A}_\alpha\|<\lambda$, ${\frak A}_\alpha$ increasingly continuous, $\langle {\frak A}_\beta:\beta\le\alpha\rangle\in {\frak A}_{\alpha+1}$, ${\frak A}_\alpha\cap\lambda$ is an ordinal, ${\frak A} =\bigcup\limits_{\alpha<\lambda}{\frak A}_\alpha$ and \[E=:\{\alpha:{\frak A}_\alpha\cap\lambda=\alpha\}.\] Note: $\prod\bar m\subseteq {\frak A}$ (as $\prod\bar m\in {\frak A}$ and $|\prod\bar m|=2^{\aleph_0}$). \noindent NOTE: By ${\bf U}^{<\mu}_{J^2_{\bar m}}(\lambda)=\lambda$, \begin{description} \item[$(*)$] \underbar{if} $x_\eta\in{\rm lev}_\omega(T^*)$ for $\eta\in\prod\bar m$ \underbar{then} for some $A\in (J^2_{\bar m})^+$ the set $\langle(\eta, x_\eta):\eta\in A\rangle$ belongs to ${\frak A}$. But then for some $\nu\in \bigcup\limits_k\prod\limits_{i<k} m_i$, the set $A$ is dense above $\nu$ (by the definition of $J^2_{\bar m}$) and hence: if the mapping $\eta\mapsto x_\eta$ is continuous then $\langle x_\rho:\nu\triangleleft\rho\in\prod\bar m \rangle\in {\frak A}$. \end{description} For $\delta\in S$ such that $C_\delta \subseteq E$ we let \[\begin{ALIGN} P^0_\delta=P^0_\delta({\frak A})=\biggl\{\bar x:&\bar x=\langle x_\rho:\rho \in t\rangle\in {\frak A}\mbox{ and } x_\rho\in{\rm lev}_{\ell g(\rho)}T^*,\\ &\mbox{the mapping }\rho\mapsto x_\rho\mbox{ preserves all of the relations:}\\ &\ell g(\rho)=n,\rho_1\triangleleft\rho_2, \neg(\rho_1\triangleleft\rho_2), \neg(\rho_1=\rho_2),\\ &\rho_1\cap\rho_2=\rho_3\mbox{ (and so }\ell g(\rho_1\cap\rho_2)=n\mbox{ is preserved});\\ &\mbox{and } t\subseteq\bigcup\limits_{\alpha\le\omega}\prod\limits_{i< \alpha} m_i\biggr\}. \end{ALIGN}\] Assume $\bar x=\langle x_\rho:\rho\in t\rangle\in P^0_\delta$. Let \[{\rm inv}(\bar x,C_\delta,T^*,{\bar {\frak A}})=:\big\{\alpha\in C_\delta: (\exists\rho\in{\rm Dom}(\bar x))(x_\rho\in {\frak A}_{\min(C_\delta \setminus (\alpha+1))}\setminus {\frak A}_\alpha\big)\}.\] Let \[\begin{array}{l} {\rm Inv}(C_\delta,T^*,\bar{\frak A})=:\\ \big\{a:\mbox{for some }\bar x\in P^0_\delta,\ \ a\mbox{ is a countable subset of }{\rm inv}(\bar x,C_0,T^*,\bar{\frak A})\big\}. \end{array}\] Note: ${\rm inv}(\bar x,C_0,T^*,\bar{\frak A})$ has cardinality at most continuum, so ${\rm Inv}(C_0,T^*,\bar{\frak A})$ is a family of $\le 2^{\aleph_0}\times |{\frak A}|=\lambda$ countable subsets of $C_\delta$. We continue as before. Let $\alpha_{\delta,\varepsilon}$ be the $\varepsilon$-th member of $C_\delta$ for $\varepsilon<\sum\limits_{n<\omega} \lambda_n$. So as $\lambda< \mu^{\aleph_0}$,$\mu> 2^{\aleph_0}$ clearly $\lambda< {\rm cf}([\lambda]^{\aleph_0}, \subseteq)$ (equivalently $\lambda<{\rm cov}(\mu,\mu,\aleph_1,2)$) hence we can find $\gamma_n\in (\bigcup\limits_{\ell<n}\lambda_\ell,\lambda_n)$ limit so such that for each $\delta\in S$, $a\in{\rm Inv}(C_\delta,T^*,\bar{\frak A})$ we have $\{\gamma_n+\ell:n<\omega\mbox{ and }\ell<m_i\}\cap a$ is bounded in $\mu$. Now we can find $T$ such that ${\rm lev}_n(T)=\prod\limits_{\ell<n}\lambda_\ell$ and \[\begin{ALIGN} {\rm lev}_\omega(T)=\big\{\bar\beta: &\bar\beta=\langle\beta_\ell:\ell<\omega \rangle,\mbox{ and for some }\delta \in S,\mbox{ for every }\ell<\omega\\ &\mbox{we have }\gamma'_\ell\in\{\alpha_{\delta,\gamma_\ell+m}:m<m_i\} \big\}. \end{ALIGN} \] So, if $T^*$ is universal there is an embedding $f:T\longrightarrow T^*$, and hence \[E'=\{\alpha\in E:{\frak A}_\alpha\mbox{ is closed under } f\mbox{ and } f^{-1}\}\] is a club of $\lambda$. By the choice of $\bar C$ for some $\delta\in S$ we have $C_\delta\subseteq E'$. Now use $(*)$ with $x_\eta=f(\bar\beta^{\delta, \eta})$, where $\beta^{\delta,\eta}_\ell=\alpha_{\delta,\gamma_\ell+ \eta(\ell)}\in{\rm lev}_\omega(T)$. Thus we get $A\in (J^2_{\bar m})^+$ such that $\{(\eta,x_\eta):\eta\in A\}\in{\frak A}$, there is $\nu\in\bigcup\limits_k \prod\limits_{i<k} m_i$ such that $A$ is dense above $\nu$, hence as $f$ is continuous, $\langle(\eta,x_\eta):\nu\triangleleft\eta\in\prod\bar m\rangle \in {\frak A}$. So $\langle x_\eta:\eta\in\prod\bar m,\nu\trianglelefteq\eta \rangle\in P^0_\delta({\frak A})$, and hence the set \[\{\alpha_{\delta,\gamma_\ell+m}:\ell\in [\ell g(\nu),\omega)\mbox{ and } m< m_\ell\}\cup\{\alpha_{\delta,\gamma_i+\nu(i)}:\ell<\ell g(\nu)\}\] is ${\rm inv}(\bar x,C_\delta,T^*,{\frak A})$. Hence \[a=\{\alpha_{\delta,\gamma_\ell}\!:\ell\in [\ell g(\nu),\omega)\}\in{\rm Inv}( C_\delta,T^*,{\frak A}),\] contradicting ``$\{\alpha_{\delta,\gamma_\ell}:\ell<\omega\}$ has finite intersection with any $a\in{\rm Inv}(C_\delta,T^*,{\frak A})$''. \begin{Remark} \label{7.3} We can a priori fix a set of $\aleph_0$ candidates and say more on their order of appearance, so that ${\rm Inv}(\bar x,C_\delta,T^*,{\bar{\frak A}})$ has order type $\omega$. This makes it easier to phrase a true invariant, i.e. $\langle (\eta_n,t_n):n<\omega\rangle$ is as above, $\langle\eta_n:n<\omega\rangle$ lists ${}^{\omega >}\omega$ with no repetition, $\langle t_n\cap {}^\omega \omega:n<\omega\rangle$ are pairwise disjoint. If $x_\rho\in{\rm lev}_\omega(T^*)$ for $\rho\in {}^\omega\omega$, $\bar T^*=\langle\bar T^*_\zeta:\zeta<\lambda \rangle$ representation \[\begin{array}{l} {\rm inv}(\langle x_\rho:\rho\in {}^\omega\omega\rangle,C_\delta,\bar T^*)=\\ \big\{\alpha\in C_\delta:\mbox{for some } n,\ (\forall\rho)[\rho\in t_n\cap {}^\omega\omega\quad\Rightarrow\quad x_\rho\in T^*_{\min(C_\delta\setminus (\alpha+1))}\setminus T^*_\alpha]\big\}. \end{array}\] \end{Remark} \begin{Remark} \label{7.4} If we have $\Gamma\in (J^2_{\bar m})^+$, $\Gamma$ non-meagre, $J=J^2_m \restriction\Gamma$ and ${\bf U}^2_J(\lambda)<\lambda^{\aleph_0}$ then we can weaken the cardinal assumptions to: \[\bar\lambda=\langle\lambda_n:n<\omega\rangle\char 94\langle\lambda\rangle, \qquad \mu=\sum_n\lambda_n,\qquad\lambda_n<\lambda_{n+1},\] \[\mu^+<\lambda={\rm cf}(\lambda)\qquad\mbox{ and }\qquad {\bf U}^2_J(\lambda)<{\rm cov}(\mu, \mu,\aleph_1,2) (\mbox{see }\ref{0.4}).\] The proof is similar. \end{Remark} \section{Universals in singular cardinals} In \S3, \S5, \ref{7.2}, we can in fact deal with ``many'' singular cardinals $\lambda$. This is done by proving a stronger assertion on some regular $\lambda$. Here ${\frak K}$ is a class of models. \begin{Lemma} \label{8.1} \begin{enumerate} \item There is no universal member in ${\frak K}_{\mu^*}$ if for some $\lambda <\mu^*$, $\theta\ge 1$ we have: \begin{description} \item[\mbox{$\otimes_{\lambda,\mu^*,\theta}[{\frak K}]$}] not only there is no universal member in ${\frak K}_\lambda$ but if we assume: \[\langle M_i:i<\theta\rangle\ \mbox{ is given, }\ \|M_i\|\le\mu^*<\prod_n \lambda_n,\ M_i\in {\frak K},\] then there is a structure $M$ from ${\frak K}_\lambda$ (in some cases of a simple form) not embeddable in any $M_i$. \end{description} \item Assume \begin{description} \item[$\otimes^\sigma_1$] $\langle\lambda_n:n<\omega\rangle$ is given, $\lambda^{\aleph_0}_n<\lambda_{n+1}$, \[\mu=\sum_{n<\omega}\lambda_n<\lambda={\rm cf}(\lambda)\leq \mu^*<\prod_{n<\omega}\lambda_n\] and $\mu^+<\lambda$ or at least there is a club guessing $\bar C$ as in $(**)^1_\lambda$ (ii) of \ref{3.4} for $(\lambda,\mu)$. \end{description} \underbar{Then} there is no universal member in ${\frak K}_{\mu^*}$ (and moreover $\otimes_{\lambda,\mu^*,\theta}[{\frak K}]$ holds) in the following cases \begin{description} \item[$\otimes_2$(a)] for torsion free groups, i.e. ${\frak K}={\frak K}^{rtf}_{\bar\lambda}$ if ${\rm cov}(\mu^*,\lambda^+,\lambda^+,\lambda)< \prod\limits_{n<\omega}\lambda_n$, see notation \ref{0.4} on ${\rm cov}$) \item[\quad(b)] for ${\frak K}={\frak K}^{tcf}_{\bar\lambda}$, \item[\quad(c)] for ${\frak K}={\frak K}^{tr}_{\bar\lambda}$ as in \ref{7.2} - ${\rm cov}({\bf U}_{J^3_{\bar m}}(\mu^*),\lambda^+,\lambda^+,\lambda)<\prod\limits_{n< \omega}\lambda_n$, \item[(d)] for ${\frak K}^{rs(p)}_{\bar\lambda}$: like case (c) (for appropriate ideals), replacing $tr$ by $rs(p)$. \end{description} \end{enumerate} \end{Lemma} \begin{Remark} \label{8.1A} \begin{enumerate} \item For \ref{7.2} as $\bar m=\langle\omega:i<\omega\rangle$ it is clear that the subtrees $t_n$ are isomorphic. We can use $m_i\in [2,\omega)$, and use coding; anyhow it is immaterial since ${}^\omega \omega,{}^\omega 2$ are similar. \item We can also vary $\bar\lambda$ in \ref{8.1} $\otimes_2$, case (c). \item We can replace ${\rm cov}$ in $\otimes_2$(a),(c) by \[\sup{\rm pp}_{\Gamma(\lambda)}(\chi):{\rm cf}(\chi)=\lambda,\lambda<\chi\le {\bf U}_{J^3_{\bar m}}(\mu^*)\}\] (see \cite[5.4]{Sh:355}, \ref{2.3}). \end{enumerate} \end{Remark} \proof Should be clear, e.g.\\ {\em Proof of Part 2), Case (c)}\ \ \ Let $\langle T_i:i<i^* \rangle$ be given, $i^*<\prod\limits_{n<\omega}\lambda_n$ such that \[\|T_i\|\le\mu^*\quad\mbox{ and }\quad\mu^\otimes=:{\rm cov}({\bf U}_{J^3_{\bar m}}(\mu^*),\lambda^+,\lambda^+,\lambda)<\prod_{n<\omega}\lambda_n.\] By \cite[5.4]{Sh:355} and ${\rm pp}$ calculus (\cite[2.3]{Sh:355}), $\mu^\otimes= {\rm cov}(\mu^\otimes,\lambda^+,\lambda^+,\lambda)$. Let $\chi=\beth_7(\lambda)^+$. For $i<i^*$ choose ${\frak B}_i\prec (H(\chi)\in <^*_\chi)$, $\|{\frak B}_i\|=\mu^\otimes$, $T_i\in {\frak B}_i$, $\mu^\otimes +1\subseteq {\frak B}_i$. Let $\langle Y_\alpha: \alpha<\mu^\otimes\rangle$ be a family of subsets of $T_i$ exemplifying the Definition of $\mu^\otimes= {\rm cov}(\mu^\otimes,\lambda^+,\lambda^+,\lambda)$.\\ Given $\bar x=\langle x_\eta:\eta\in {}^\omega\omega\rangle$, $x_\eta\in {\rm lev}_\omega(T_i)$, $\eta\mapsto x_\eta$ continuous (in our case this means $\ell g(\eta_1\cap\eta_2)=\ell g(x_{\eta_1}\cap x_{\eta_2})=:\ell g(\max \{\rho:\rho\triangleleft\eta_1\ \&\ \rho\triangleleft\eta_2\})$. Then for some $\eta\in {}^{\omega>}\omega$, \[\langle x_\rho:\eta\triangleleft\rho\in {}^\omega\omega\rangle\in{\frak B}.\] So given $\left<\langle x^\zeta_\eta:\eta\in {}^\omega\omega\rangle:\zeta< \lambda\right>$, $x^\zeta_\eta\in{\rm lev}_\omega(T_i)$ we can find $\langle (\alpha_j,\eta_j):j<j^*<\lambda\rangle$ such that: \[\bigwedge_{\zeta<\lambda}\bigvee_j\langle x^\zeta_\eta:\eta_j\triangleleft \eta\in {}^\omega\omega\rangle\in Y_\alpha.\] Closing $Y_\alpha$ enough we can continue as usual. \hfill$\square_{\ref{8.1}}$ \section{Metric spaces and implications} \begin{Definition} \label{9.1} \begin{enumerate} \item ${\frak K}^{mt}$ is the class of metric spaces $M$ (i.e. $M=(|M|,d)$, $|M|$ is the set of elements, $d$ is the metric, i.e. a two-place function from $|M|$ to ${\Bbb R}^{\geq 0}$ such that $d(x,y)=0\quad\Leftrightarrow\quad x=0$ and $d(x,z)\le d(x,y)+d(y,z)$ and $d(x,y) = d(y,x)$). An embedding $f$ of $M$ into $N$ is a one-to-one function from $|M|$ into $|N|$ which is continuous, i.e. such that: \begin{quotation} \noindent if in $M$, $\langle x_n:n<\omega\rangle$ converges to $x$ \noindent then in $N$, $\langle f(x_n):n<\omega\rangle$ converges to $f(x)$. \end{quotation} \item ${\frak K}^{ms}$ is defined similarly but ${\rm Rang}(d)\subseteq\{2^{-n}:n <\omega\}\cup\{0\}$ and instead of the triangular inequality we require \[d(x,y)=2^{-i},\qquad d(y,z)=2^{-j}\quad \Rightarrow\quad d(x,z) \le 2^{-\min\{i-1,j-1\}}.\] \item ${\frak K}^{tr[\omega]}$ is like ${\frak K}^{tr}$ but $P^M_\omega=|M|$ and embeddings preserve $x\;E_n\;y$ (not necessarily its negation) are one-to-one, and remember $\bigwedge\limits_n x\;E_n\;y\quad\Rightarrow\quad x \restriction n = y \restriction n$). \item ${\frak K}^{mt(c)}$ is the class of semi-metric spaces $M=(|M|,d)$, which means that for the constant $c\in\Bbb R^+$ the triangular inequality is weakened to $d(x,z)\le cd(x,y)+cd(y,z)$ with embedding as in \ref{9.1}(1) (so for $c=1$ we get ${\frak K}^{mt}$). \item ${\frak K}^{mt[c]}$ is the class of pairs $(A,d)$ such that $A$ is a non-empty set, $d$ a two-place symmetric function from $A$ to ${\Bbb R}^{\ge 0}$ such that $[d(x,y)=0\ \Leftrightarrow\ x=y]$ and \[d(x_0,x_n)\le c\sum\limits_{\ell<n} d(x_\ell,x_{\ell+1})\ \ \mbox{ for any $n<\omega$ and $x_0,\ldots,x_n\in A$.}\] \item ${\frak K}^{ms(c)}$, ${\frak K}^{ms[c]}$ are defined parallely. \item ${\frak K}^{rs(p),\mbox{pure}}$ is defined like ${\frak K}^{rs(p)}$ but the embeddings are pure. \end{enumerate} \end{Definition} \begin{Remark} There are, of course, other notions of embeddings; isometric embeddings if $d$ is preserved, co-embeddings if the image of an open set is open, bi-continuous means an embedding which is a co-embedding. The isometric embedding is the weakest, its case is essentially equivalent to the ${\frak K}^{tr}_{\lambda}$ case (as in \ref{9.3}(3)); for the open case there is a universal: discrete space. The universal for ${\frak K}^{mt}_\lambda$ under bicontinuous case exist in cardinality $\lambda^{\aleph_0}$, see \cite{Ko57}. \end{Remark} \begin{Definition} \label{9.1A} \begin{enumerate} \item ${\rm Univ}^0({\frak K}^1,{\frak K}^2)=\{(\lambda,\kappa,\theta):$ there are $M_i\in {\frak K}^2_\kappa$ for $i<\theta$ such that any $M\in {\frak K}^1_\lambda$ can be embedded into some $M_i\}$. We may omit $\theta$ if it is 1. We may omit the superscript 0. \item ${\rm Univ}^1({\frak K}^1,{\frak K}^2)=\{(\lambda,\kappa,\theta):$ there are $M_i\in {\frak K}^2_\kappa$ for $i<\theta$ such that any $M\in {\frak K}^1_\lambda$ can be represented as the union of $<\lambda$ sets $A_\zeta$ ($\zeta<\zeta^*<\lambda)$ such that each $M\restriction A_\zeta$ can be embedded into some $M_i\}$ and is a $\leq_{{\frak K}^1}$-submodel of $M$. \item If above ${\frak K}^1={\frak K}^2$ we write it just once; (naturally we usually assume ${\frak K}^1 \subseteq {\frak K}^2$). \end{enumerate} \end{Definition} \begin{Remark} \begin{enumerate} \item We prove our theorems for $Univ^0$, we can get parallel things for ${\rm Univ}^1$. \item Many previous results of this paper can be rephrased using a pair of classes. \item We can make \ref{9.2} below deal with pairs and/or function $H$ changing cardinality. \item ${\rm Univ}^\ell$ has the obvious monotonicity properties. \end{enumerate} \end{Remark} \begin{Proposition} \label{9.2} \begin{enumerate} \item Assume ${\frak K}^1,{\frak K}^2$ has the same models as their members and every embedding for ${\frak K}^2$ is an embedding for ${\frak K}^1$.\\ Then ${\rm Univ}({\frak K}^2)\subseteq{\rm Univ}({\frak K}^1)$. \item Assume there is for $\ell=1,2$ a function $H_\ell$ from ${\frak K}^\ell$ into ${\frak K}^{3-\ell}$ such that: \begin{description} \item[(a)] $\|H_1(M_1)\|=\|M_1\|$ for $M_1\in {\frak K}^1$, \item[(b)] $\|H_2(M_2)\|=\|M_2\|$ for $M_2\in {\frak K}^2$, \item[(c)] if $M_1\in {\frak K}^1$, $M_2\in {\frak K}^2$, $H_1(M_1)\in {\frak K}^2$ is embeddable into $M_2$ \underbar{then} $M_1$ is embeddable into $H_2(M_2)\in {\frak K}^1$. \end{description} \underbar{Then}\quad ${\rm Univ}({\frak K}^2)\subseteq{\rm Univ}({\frak K}^1)$. \end{enumerate} \end{Proposition} \begin{Definition} \label{9.2A} We say ${\frak K}^1\le {\frak K}^2$ if the assumptions of \ref{9.2}(2) hold. We say ${\frak K}^1\equiv {\frak K}^2$ if ${\frak K}^1\le {\frak K}^2 \le {\frak K}^1$ (so larger means with fewer cases of universality). \end{Definition} \begin{Theorem} \label{9.3} \begin{enumerate} \item The relation ``${\frak K}^1\le {\frak K}^2$'' is a quasi-order (i.e. transitive and reflexive). \item If $({\frak K}^1,{\frak K}^2)$ are as in \ref{9.2}(1) then ${\frak K}^1 \le {\frak K}^2$ (use $H_1 = H_2 =$ the identity). \item For $c_1>1$ we have ${\frak K}^{mt(c_1)}\equiv {\frak K}^{mt[c_1]}\equiv {\frak K}^{ms[c_1]}\equiv {\frak K}^{ms(c_1)]}$. \item ${\frak K}^{tr[\omega]} \le {\frak K}^{rs(p)}$. \item ${\frak K}^{tr[\omega]} \le {\frak K}^{tr(\omega)}$. \item ${\frak K}^{tr(\omega)} \le {\frak K}^{rs(p),\mbox{pure}}$. \end{enumerate} \end{Theorem} \proof 1)\ \ Check.\\ 2)\ \ Check. \\ 3)\ \ Choose $n(*)<\omega$ large enough and ${\frak K}^1,{\frak K}^2$ any two of the four. We define $H_1,H_2$ as follows. $H_1$ is the identity. For $(A,d) \in{\frak K}^\ell$ let $H_\ell((A,d))=(A,d^{[\ell]})$ where $d^{[\ell]}(x,y)= \inf\{1/(n+n(*)):2^{-n}\ge d(x,y)\}$ (the result is not necessarily a metric space, $n(*)$ is chosen so that the semi-metric inequality holds). The point is to check clause (c) of \ref{9.2}(2); so assume $f$ is a function which ${\frak K}^2$-embeds $H_1((A_1,d_1))$ into $(A_2,d_2)$; but \[H_1((A_1,d_1))=(A_1,d_1),\quad H_2((A_2,d_2))=(A_2,d^{[2]}_2),\] so it is enough to check that $f$ is a function which ${\frak K}^1$-embeds $(A_1,d^{[1]}_1)$ into $(A_2,d^{[2]}_2)$ i.e. it is one-to-one (obvious) and preserves limit (check).\\ 4)\ \ For $M=(A,E_n)_{n<\omega}\in {\frak K}^{tr[\omega]}$, without loss of generality $A\subseteq {}^\omega\lambda$ and \[\eta E_n\nu\qquad\Leftrightarrow\qquad\eta\in A\ \&\ \nu\in A\ \&\ \eta\restriction n=\nu\restriction n.\] Let $B^+=\{\eta\restriction n:\eta\in A\mbox{ and } n<\omega\}$. We define $H_1(M)$ as the (Abelian) group generated by \[\{x_\eta:\eta\in A\cup B\}\cup\{y_{\eta,n}:\eta\in A,n<\omega\}\] freely except \[\begin{array}{rcl} p^{n+1}x_\eta=0 &\quad\mbox{\underbar{if}}\quad &\eta\in B, \ell g(\eta)=n\\ y_{\eta,0}=x_\eta &\quad\mbox{\underbar{if}}\quad &\eta\in A\\ py_{\eta,n+1}-y_\eta=x_{\eta\restriction n} &\quad\mbox{\underbar{if}}\quad &\eta\in A, n<\omega\\ p^{n+1}y_{\eta, n}=0 &\quad\mbox{\underbar{if}}\quad &\eta\in B, n<\omega. \end{array}\] For $G\in {\frak K}^{rs(p)}$ let $H_2(G)$ be $(A,E_n)_{n<\omega}$ with: \[A = G,\quad xE_ny\quad\underbar{iff}\quad G\models\mbox{``}p^n\mbox{ divides }(x-y)\mbox{''.}\] $H_2(G)\in {\frak K}^{tr[\omega]}$ as ``$G$ is separable" implies $(\forall x)(x \ne 0\ \Rightarrow\ (\exists n)[x\notin p^nG])$. Clearly clauses (a), (b) of Definition \ref{9.1}(2) hold. As for clause (c), assume $(A,E_n)_{n<\omega} \in {\frak K}^{tr[\omega]}$. As only the isomorphism type counts without loss of generality $A\subseteq {}^\omega\lambda$. Let $B=\{\eta\restriction n:n< \omega:\eta\in A\}$ and $G=H_1((A,E_n)_{n<\omega})$ be as above. Suppose that $f$ embeds $G$ into some $G^*\in {\frak K}^{rs(p)}$, and let $(A^*,E^*_n)_{n< \omega}$ be $H_2(G^*)$. We should prove that $(A,E_n)_{n<\omega}$ is embeddable into $(A^*,E^*_n)$.\\ Let $f^*:A\longrightarrow A^*$ be $f^*(\eta)=x_\eta\in A^*$. Clearly $f^*$ is one to one from $A$ to $A^*$; if $\eta E_n \nu$ then $\eta\restriction n=\nu \restriction n$ hence $G \models p^n \restriction (x_\eta-x_\nu)$ hence $(A^*,A^*_n)_{n<\omega}\models\eta E^*_n\nu$. \hfill$\square_{\ref{9.3}}$ \begin{Remark} \label{9.3A} In \ref{9.3}(4) we can prove ${\frak K}^{tr[\omega]}_{\bar\lambda}\le{\frak K}^{rs(p)}_{\bar\lambda}$. \end{Remark} \begin{Theorem} \label{9.4} \begin{enumerate} \item ${\frak K}^{mt} \equiv {\frak K}^{mt(c)}$ for $c \ge 1$. \item ${\frak K}^{mt} \equiv {\frak K}^{ms[c]}$ for $c > 1$. \end{enumerate} \end{Theorem} \proof 1)\ \ Let $H_1:{\frak K}^{mt}\longrightarrow {\frak K}^{mt(c)}$ be the identity. Let $H_2:{\frak K}^{mt(c)}\longrightarrow {\frak K}^{mt}$ be defined as follows:\\ $H_2((A,d))=(A,d^{mt})$, where \[\begin{array}{l} d^{mt}(y,z)=\\ \inf\big\{\sum\limits^n_{\ell=0} d(x_\ell,x_{\ell,n}):n<\omega\ \&\ x_\ell\in A\mbox{ (for $\ell\le n$)}\ \&\ x_0=y\ \&\ x_n=z\big\}. \end{array}\] Now \begin{description} \item[$(*)_1$] $d^{mt}$ is a two-place function from $A$ to ${\Bbb R}^{\ge 0}$, is symmetric, $d^{mt}(x,x)=0$ and it satisfies the triangular inequality. \end{description} This is true even on ${\frak K}^{mt(c)}$, but here also \begin{description} \item[$(*)_2$] $d^{mt}(x,y) = 0 \Leftrightarrow x=0$. \end{description} [Why? As by the Definition of ${\frak K}^{mt[c]},d^{mt}(x,y)\ge{\frac 1c} d(x,y)$. Clearly clauses (a), (b) of \ref{9.2}(2) hold.]\\ Next, \begin{description} \item[$(*)_3$] If $M_1,N\in {\frak K}^{mt}$, $f$ is an embedding (for ${\frak K}^{mt}$) of $M_1$ into $N$ then $f$ is an embedding (for ${\frak K}^{mt[c]}$) of $H_1(M)$ into $H_1(N)$ \end{description} [why? as $H_1(M)=M$ and $H_2(N)=N$], \begin{description} \item[$(*)_4$] If $M,N\in {\frak K}^{mt[c]}$, $f$ is an embedding (for ${\frak K}^{mt[c]}$) of $M$ into $N$ then $f$ is an embedding (for ${\frak K}^{mt}$) of $H_2(M)$ into $H_1(M)$ \end{description} [why? as $H^*_\ell$ preserves $\lim\limits_{n\to\infty} x_n=x$ and $\lim\limits_{n\to\infty} x_n\ne x$]. So two applications of \ref{9.2} give the equivalence. \\ 2)\ \ We combine $H_2$ from the proof of (1) and the proof of \ref{9.3}(3). \hfill$\square_{\ref{9.4}}$ \begin{Definition} \label{9.6} \begin{enumerate} \item If $\bigwedge\limits_n\mu_n=\aleph_0$ let \[\hspace{-0.5cm}\begin{array}{ll} J^{mt}=J^{mt}_{\bar\mu}=\big\{A\subseteq\prod\limits_{n<\omega}\mu_n:& \mbox{for every } n \mbox{ large enough, } \\ \ & \mbox{ for every }\eta\in\prod\limits_{\ell <n}\mu_\ell\\ \ &\mbox{the set }\{\eta'(n):\eta\triangleleft\eta'\in A\}\mbox{ is finite} \big\}. \end{array}\] \item Let $T=\bigcup\limits_{\alpha\le\omega}\prod\limits_{n<\alpha}\mu_n$, $(T,d^*)$ be a metric space such that \[\prod_{\ell<n}\mu_\ell\cap\mbox{closure}\left(\bigcup_{m<n}\prod_{\ell<m} \mu_\ell\right)=\emptyset;\] now \[\begin{ALIGN} I^{mt}_{(T,d^*)}=:\big\{A\subseteq\prod\limits_{n<\omega}\mu_n:&\mbox{ for some }n,\mbox{ the closure of } A\mbox{ (in $(T,d^*)$)}\\ &\mbox{ is disjoint to }\bigcup\limits_{m\in [n,\omega)}\prod\limits_{\ell <m} \mu_\ell\big\}. \end{ALIGN}\] \item Let $H\in {\frak K}^{rs(p)}$, $\bar H=\langle H_n:n<\omega\rangle$, $H_n \subseteq H$ pure and closed, $n<m\ \Rightarrow\ H_n\subseteq H_m$ and $\bigcup\limits_{n<\omega} H_n$ is dense in $H$. Let \[\begin{array}{ll} I^{rs(p)}_{H,\bar H}=:\big\{A\subseteq H:&\mbox{for some } n\mbox{ the closure of }\langle A\rangle_H\mbox{ intersected with}\\ &\bigcup\limits_{\ell<\omega}H_\ell\mbox{ is included in }H_n\big\}. \end{array}\] \end{enumerate} \end{Definition} \begin{Proposition} \label{9.5} Suppose that $2^{\aleph_0}<\mu$ and $\mu^+<\lambda={\rm cf}(\lambda)< \mu^{\aleph_0}$ and \begin{description} \item[$(*)_\lambda$] ${\bf U}_{J^{mt}_{\bar\mu}}(\lambda)=\lambda$ or at least ${\bf U}_{J^{mt}_{\bar\mu}}(\lambda)<\lambda^{\aleph_0}$ for some $\bar\mu=\langle \mu_n:n<\omega\rangle$ such that $\prod\limits_{n<\omega}\mu_n<\lambda$. \end{description} Then ${\frak K}^{mt}_\lambda$ has no universal member. \end{Proposition} \begin{Proposition} \label{9.7} \begin{enumerate} \item $J^{mt}$ is $\aleph_1$-based. \item The minimal cardinality of a set which is not in the $\sigma$-ideal generated by $J^{mt}$ is ${\frak b}$. \item $I^{mt}_{(T,d^*)},I^{rs(p)}_{H,\bar H}$ are $\aleph_1$-based. \item $J^{mt}$ is a particular case of $I^{mt}_{(T,d^*)}$ (i.e. for some choice of $(T,d^*)$). \item $I^0_{\bar \mu}$ is a particular case of $I^{rs(p)}_{H,\bar H}$. \end{enumerate} \end{Proposition} \proof of \ref{9.5}. Let $$ \begin{array}{ll} T_\alpha=\{(\eta, \nu)\in{}^\alpha\lambda\times {}^\alpha(\omega+ 1):& \mbox{ for every }n\mbox{ such that }n+1< \alpha\\ \ & \mbox{ we have }\nu(n)< \omega\} \end{array} $$ and for $\alpha\le\omega$ let $T=\bigcup\limits_{\alpha\le \omega}T_\alpha$. We define on $T$ the relation $<_T$: \[(\eta_1,\nu_1)\le(\eta_1,\nu_2)\quad\mbox{ iff }\quad\eta_1\trianglelefteq \eta_2\ \&\ \nu_1\triangleleft\nu_2.\] We define a metric:\\ if $(\eta_1,\nu_1)\ne(\eta_2,\nu_2)\in T$ and $(\eta,\nu)$ is their maximal common initial segment and $(\eta,\nu)\in T$ then necessarily $\alpha= \ell g((\eta,\nu))<\omega$ and we let: \begin{quotation} \noindent if $\eta_1(\alpha)\ne\eta_2(\alpha)$ then \[d\left((\eta_1,\nu_1),(\eta_2,\nu_2)\right)=2^{-\sum\{\nu(\ell):\ell< \alpha\}},\] if $\eta_1(\alpha)=\eta_2(\alpha)$ (so $\nu_1(\alpha)\ne\nu_2(\alpha)$ then \[d\left((\eta_1,\nu_1),(\eta_2,\nu_2)\right)=2^{-\sum\{\nu(\ell):\ell<\alpha \}}\times 2^{-\min\{\nu_1(\alpha),\nu_2(\alpha)\}}.\] \end{quotation} Now, for every $S\subseteq\{\delta<\lambda:{\rm cf}(\delta)=\aleph_0\}$, and $\bar \eta=\langle\eta_\delta:\delta\in S\rangle$, $\eta_\delta\in {}^\omega \delta$, $\eta_\delta$ increasing let $M_\eta$ be $(T,d)\restriction A_{\bar \eta}$, where \[A_{\bar\eta}=\bigcup_{n<\omega} T_n\cup\{(\eta_\delta,\nu):\delta\in S,\;\nu \in {}^\omega\omega\}.\] The rest is as in previous cases (note that $\langle(\eta\char 94\langle \alpha \rangle,\nu\char 94\langle n\rangle):n<\omega\rangle$ converges to $(\eta\char 94\langle\alpha\rangle,\nu\char 94\langle\omega\rangle)$ and even if $(\eta\char 94\langle \alpha\rangle, \nu\char 94\langle n\rangle)\leq (\eta_n, \nu_n)\in T_\omega$ then $\langle(\eta_n, \nu_n): n<\omega\rangle$ converge to $(\eta\char 94 \langle \alpha\rangle, \nu\char 94\langle \omega\rangle)$). \hfill$\square_{\ref{9.7}}$ \begin{Proposition} \label{9.8} If ${\rm IND}_{\chi'}(\langle\mu_n:n<\omega\rangle)$, then $\prod\limits_{n<\omega} \mu_n$ is not the union of $\le\chi$ members of $I^0_{\bar\mu}$ (see Definition \ref{5.5A} and Theorem \ref{5.5}). \end{Proposition} \proof Suppose that $A_\zeta=\{\sum\limits_{n<\omega} p^nx^n_{\alpha_n}:\langle \alpha_n:n<\omega\rangle\in X_\zeta\}$ and $\alpha_n<\mu_n$ are such that if $\sum p^nx^n_{\alpha_n}\in A_\zeta$ then for infinitely many $n$ for every $k<\omega$ there is $\langle \beta_n:n<\omega\rangle$, \[(\forall\ell<k)[\alpha_\ell=\beta_\ell\ \ \Leftrightarrow\ \ \ell=n]\qquad \mbox{ and }\qquad\sum_{n<\omega}p^nx^n_{\beta_n}\in A_\zeta\ \ \mbox{ (see \S5).}\] This clearly follows. \hfill$\square_{\ref{9.8}}$ \section{On Modules} Here we present the straight generalization of the one prime case like Abelian reduced separable $p$-groups. This will be expanded in \cite{Sh:622} (including the proof of \ref{10.4new}). \begin{Hypothesis} \label{10.1} \begin{description} \item[(A)] $R$ is a ring, $\bar{\frak e}=\langle{\frak e}_n:n<\omega\rangle$, ${\frak e}_n$ is a definition of an additive subgroup of $R$-modules by an existential positive formula (finitary or infinitary) decreasing with $n$, we write ${\frak e}_n(M)$ for this additive subgroup, ${\frak e}_\omega(M)= \bigcap\limits_n {\frak e}_n(M)$. \item[(B)] ${\frak K}$ is the class of $R$-modules. \item[(C)] ${\frak K}^*\subseteq {\frak K}$ is a class of $R$-modules, which is closed under direct summand, direct limit and for which there is $M^*$, $x^* \in M^*$, $M^*=\bigoplus\limits_{\ell\le n} M^*_\ell\oplus M^{**}_n$, $M^*_n \in {\frak K}$, $x^*_n\in {\frak e}_n(M^*_n)\setminus {\frak e}_{n+1}(M^*)$, $x^*-\sum\limits_{\ell<n} x^*_\ell\in {\frak e}_n(M^*)$. \end{description} \end{Hypothesis} \begin{Definition} \label{10.2} For $M_1,M_2\in {\frak K}$, we say $h$ is a $({\frak K},\bar {\frak e})$-homomorphism from $M_1$ to $M_2$ if it is a homomorphism and it maps $M_1 \setminus {\frak e}_\omega(M_1)$ into $M_2\setminus {\frak e}_\omega(M_2)$; we say $h$ is an $\bar {\frak e}$-pure homomorphism if for each $n$ it maps $M_1\setminus {\frak e}_n(M_1)$ into $M_2\setminus {\frak e}_n(M_2)$. \end{Definition} \begin{Definition} \label{10.3} \begin{enumerate} \item Let $H_n \subseteq H_{n+1} \subseteq H$, $\bar H=\langle H_n:n<\omega \rangle$, $c\ell$ is a closure operation on $H$, $c\ell$ is a function from ${\cal P}(H)$ to itself and \[X \subseteq c \ell(X) = c \ell(c \ell(X)).\] Define \[I_{H,\bar H,c\ell}=\big\{A\subseteq H:\mbox{for some }k<\omega\mbox{ we have } c\ell(A)\cap\bigcup_{n<\omega} H_n\subseteq H_k\big\}.\] \item We can replace $\omega$ by any regular $\kappa$ (so $H=\langle H_i:i< \kappa\rangle$). \end{enumerate} \end{Definition} \begin{Claim} \label{10.4new} Assume $|R|+\mu^+< \lambda = {\rm cf}(\lambda)< \mu^{\aleph_0}$, then for every $M\in {\frak K}_\lambda$ there is $N\in {\frak K}_\lambda$ with no $\bar {\frak e}$-pure homomorphism from $N$ into $M$. \end{Claim} \begin{Remark} In the interesting cases $c\ell$ has infinitary character.\\ The applications here are for $\kappa=\omega$. For the theory, $pcf$ is nicer for higher $\kappa$. \end{Remark} \section{Open problems} \begin{Problem} \begin{enumerate} \item If $\mu^{\aleph_0}\ge\lambda$ then any $(A,d)\in {\frak K}^{mt}_\lambda$ can be embedded into some $M'\in {\frak K}^{mt}_\lambda$ with density $\le\mu$. \item If $\mu^{\aleph_0}\ge\lambda$ then any $(A,d)\in {\frak K}^{ms}_\lambda$ can be embedded into some $M'\in {\frak K}^{ms}_\lambda$ with density $\le\mu$. \end{enumerate} \end{Problem} \begin{Problem} \begin{enumerate} \item Other inclusions on ${\rm Univ}({\frak K}^x)$ or show consistency of non inclusions (see \S9). \item Is ${\frak K}^1\le {\frak K}^2$ the right partial order? (see \S9). \item By forcing reduce consistency of ${\bf U}_{J_1}(\lambda)>\lambda+ 2^{\aleph_0}$ to that of ${\bf U}_{J_2}(\lambda)>\lambda+2^{\aleph_0}$. \end{enumerate} \end{Problem} \begin{Problem} \begin{enumerate} \item The cases with the weak ${\rm pcf}$ assumptions, can they be resolved in ZFC? (the $pcf$ problems are another matter). \item Use \cite{Sh:460}, \cite{Sh:513} to get ZFC results for large enough cardinals. \end{enumerate} \end{Problem} \begin{Problem} If $\lambda^{\aleph_0}_n<\lambda_{n+1}$, $\mu=\sum\limits_{n<\omega} \lambda_n$, $\lambda=\mu^+<\mu^{\aleph_0}$ can $(\lambda,\lambda,1)$ belong to ${\rm Univ}({\frak K})$? For ${\frak K}={\frak K}^{tr},{\frak K}^{rs(p)},{\frak K}^{trf}$? \end{Problem} \begin{Problem} \begin{enumerate} \item If $\lambda=\mu^+$, $2^{<\mu}=\lambda<2^\mu$ can $(\lambda,\lambda,1) \in {\rm Univ}({\frak K}^{\mbox{or}}=$ class of linear orders)? \item Similarly for $\lambda=\mu^+$, $\mu$ singular, strong limit, ${\rm cf}(\mu)= \aleph_0$, $\lambda<\mu^{\aleph_0}$. \item Similarly for $\lambda=\mu^+$, $\mu=2^{<\mu}=\lambda^+ <2^\mu$. \end{enumerate} \end{Problem} \begin{Problem} \begin{enumerate} \item Analyze the existence of universal member from ${\frak K}^{rs(p)}_\lambda$, $\lambda<2^{\aleph_0}$. \item \S4 for many cardinals, i.e. is it consistent that: $2^{\aleph_0}> \aleph_\omega$ and for every $\lambda< 2^{\aleph_0}$ there is a universal member of ${\frak K}^{rs(p)}_\lambda$? \end{enumerate} \end{Problem} \begin{Problem} \begin{enumerate} \item If there are $A_i\subseteq\mu$ for $i<2^{\aleph_0}$, $|A_i\cap A_j|< \aleph_0$, $2^\mu=2^{\aleph_0}$ find forcing adding $S\subseteq [{}^\omega \omega]^\mu$ universal for $\{(B, \vartriangleleft): {}^{\omega>}\omega \subseteq B\subseteq {}^{\omega\geq}\omega, |B|\leq \lambda\}$ under (level preserving) natural embedding. \end{enumerate} \end{Problem} \begin{Problem} For simple countable $T$, $\kappa=\kappa^{<\kappa}<\lambda\subseteq \kappa$ force existence of universal for $T$ in $\lambda$ still $\kappa=\kappa^{< \kappa}$ but $2^\kappa=\chi$. \end{Problem} \begin{Problem} Make \cite[\S4]{Sh:457}, \cite[\S1]{Sh:500} work for a larger class of theories more than simple. \end{Problem} See on some of these problems \cite{DjSh:614}, \cite{Sh:622}. \bibliographystyle{lit-plain}
proofpile-arXiv_065-469
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{INTRODUCTION} In Quantum Chromodynamics (QCD), quarks ($q$) and gluons ($g$) are coloured objects that carry different colour charges. Quarks have a single colour index while gluons are tensor objects carrying two colour indices. Due to this fact, quarks and gluons differ in their relative coupling strength to emit additional gluons, and, in consequence, jets originating from the fragmentation of energetic quarks and gluons are expected to show differences in their final particle multiplicities, energies and angular distributions. The masses of quarks are also fundamental parameters of the QCD lagrangian not predicted by the theory. The definition of the quark masses is however not unique because quarks are not free particles and various scenarios are possible. The perturbative pole mass, M$_q$, and the running mass, m$_q$, of the $\overline{MS}$ scheme are among the most currently used. At first order in $\alpha_s$ the predicted expression for an observable is not able to resolve the mass ambiguity and, only when second or higher order terms are included, the mass definition becomes known. At orders higher than one the renormalization scheme used as the baseline of the calculation has to be chosen and this contains the information about the mass definition. Earlier calculations of the three-jet cross section in $e^+e^-$ including mass terms already exist at O($\alpha_s$) \cite{tor,val} and have been used to evaluate mass effects for the $b$-quark when testing the universality of the strong coupling constant, $\alpha_s$. They could not however be used to evaluate the mass of the $b$-quark, m$_b$, because these calculations are ambiguous in this parameter. Recently, expressions at O($\alpha_s^2$), for the multi-jet production rate in $e^+e^-$ are available \cite{german} and, thus, they enable measuring m$_b$ in case the flavour independence of $\alpha_s$ is assumed and enough experimental precision is achieved. There are well known existing difficulties to measure all the above parameters in quantitative agreement with the predictions from perturbative QCD, since partons, quarks and gluons, are not directly observed in nature and only the stable particles, produced after the fragmentation process, are experimentally detected. However, the massive statistics and improved jet tagging techniques available at LEP presently allow overcomig these difficulties by applying restrictive selection criteria which lead to quark and gluon jet samples with high purities. The selected data samples are almost background free and small corrections to account for impurities are needed. A smaller model dependence than ever is now achieved, bringing the possibility to perform quantitative studies of quark and gluon fragmentation according to perturbative QCD. The analyses reported in here include more than 3 million Z$^0$ decays as collected by DELPHI at center-of-mass energies of $\sqrt{s}\approx$M$_Z$. In the first analysis, the ratio between the gluon jet multiplicity and the quark jet multiplicity, $r=\langle N_g \rangle /\langle N_q \rangle$, is presented and discussed in comparison with other detector results. In the second study, preliminary values of errors associated to the determination of m$_b$ at the M$_Z$ scale are given. \section{EVENT SELECTION} Gluon and quark jets were selected using hadronic three-jet events. Jets were mainly reconstructed using the {\sc Durham} algorithm although the {\sc Jade} algorithm was also used \cite{delphi_qg}, in particular, to observe the effects due to different angular particle acceptance of the various algorithms. In the gluon splitting process ($g\rightarrow q\bar{q}$), the heavy quark production is strongly suppressed \cite{gluo_qq}. Gluon jets can thus be extracted from $q\bar{q}g$ events by applying $b$ tagging techniques. The two jets which satisfy the experimental signatures of being initiated by $b$ quarks are associated to the quark jets and the remaining one is, $by$ $definition$, assigned to be the gluon jet without any further requirement. Algorithms for tagging $b$ jets exploit the fact that the decay products of long lived B hadrons have large impact parameters and/or contain inclusive high momentum leptons coming from the semileptonic decays of the B hadrons. Gluon purities of 94\% and 85\% are achieved when using these techniques, respectively. Obviously, the quark jets belonging to these events cannot be used to represent an unbiased quark sample. Thus the quark jets whose properties are to be compared with the gluon jets must be selected from other sources which in any case should preserve the same kinematics. Two possibilities have been proposed in the current literature. One consists in selecting symmetric three-jet event configurations \cite{delphi_qg,opal_qg,aleph_qg} in which one (Y) or the two (Mercedes) quark jets have similar energy to that of the gluon jet. The quark jet purities reached are $\sim$52\% and $\sim$66\%, for Y and for Mercedes events, respectively. In a second solution \cite{delphi_qg,opal_qg,l3_qqgamma} radiative $q\bar{q}\gamma$ events are selected, allowing a sample of quark jets with variable energy to be collected. In this latter case, misidentifications of $\gamma$'s due to the $\pi^\circ$ background and radiative $\tau^+\tau^-\gamma$ contamination give rise to quark jet purities of $\sim$92\%. This method gives a higher purity but unfortunately suffers from the lack of statistics. The $b$-quark purity in the $b\bar{b}g$ sample reached in the DELPHI analyses is $\sim$93\% and for the light $uds$-quarks is $\sim$80\%. Table \ref{tab:events} summarizes the number of events selected and their corresponding energy intervals. {\small \begin{table}[t] \begin{center} \caption{The three-jet event samples and their corresponding energy intervals as used in the analysis.} \begin{tabular}{lrc} \hline Event type & \# events & Jet energy range \\ \hline $q\bar{q}\gamma$ & $\begin{array}{r} 2,237 \\ \end{array}$ & 7.5~GeV - 42.5~GeV \\ $(uds)\overline{(uds)}g$ & $\begin{array}{r} 552,645 \\ \end{array}$ & 7.5~GeV - 42.5~GeV \\ $b\bar{b}g$ & $\begin{array}{r} 104,081 \\ \end{array}$ & 7.5~GeV - 42.5~GeV \\ Y & $\begin{array}{r} 74,164 \\ \end{array}$ & 19.6~GeV - 28.8~GeV \\ Mercedes & $\begin{array}{r} 9,264 \\ \end{array}$ & 27.4~GeV - 33.4~GeV \\ \hline \end{tabular} \label{tab:events} \end{center} \end{table} } \section{MULTIPLICITIES OF QUARK AND GLUON JETS} Results on the charged multiplicity of quark and gluon jets \cite{opal_qg,aleph_qg} using symmetric Y configurations and reconstructed with {\sc Durham} at 24 GeV gluon jet energy, give a ratio of $r \approx 1.23 \pm 0.04 \mbox{(stat.+syst.)}$ which does not depend on the cut-off parameter ($y_{cut}$) selected to reconstruct jets \cite{opal_qg}. It is significantly higher than one, which indicates that quark and gluons in fact fragment differently, but it remains far from the asymptotic lowest order expectation of $C_F/C_A = 9/4$, suggesting that higher order corrections and non-perturbative effects are very important to understand the measured value. A next-to-leading order correction \cite{mueller} in MLLA (Modified Leading Log Approximation) at O$(\sqrt{\alpha_s})$ already lowers the prediction towards $r$ values slightly below two and exhibits a small energy dependence due to the running of $\alpha_s$. However this is still insufficient to explain the value of $r$ determined by the experiments. Solutions based on the Monte Carlo method give a better approximation \cite{delphi_qg}. The parton shower option of the {\sc Jetset} generator \cite{jetset} which uses the Altarelli-Parisi splitting functions for the evolution of the parton shower reduces the theoretical prediction \cite{delphi_qg} for $r$. At parton level, at 24 GeV jet energy, the expected value is $\sim$1.4 and it is further reduced to $\sim$1.3 if the value of $r$ is computed after the fragmentation process. In both cases there is a clear dependence of $r$ with the jet energy \cite{delphi_qg} which can be parametrized using straight lines with slopes of ${\Delta r}/{\Delta E} = (+90 \pm 3\mbox{(stat.)})\cdot 10^{-4} \ \mbox{GeV}^{-1}$ at parton level and ${\Delta r}/{\Delta E} = (+76 \pm 2\mbox{(stat.)})\cdot 10^{-4}\ \mbox{GeV}^{-1}$ after fragmentation. The absolute value of $r$ predicted at parton level is however largely affected by the choice on the $Q_0$ parameter (cut-off at which the parton evolution stops) but has negligible influence on its relative variation with the energy, i.e., the slope. The DELPHI analysis uses symmetric and non-symmetric three-jet event configurations with quark and gluon jets of variable energy, allowing thus all these properties and predictions to be tested. A value of $r=1.23\pm0.03\ \mbox{(stat.+syst.)}$ is measured corresponding to an average jet energy of $\sim$27 GeV. The energy dependence of $r$ is also suggested at 4$\sigma$ significance level, with a fitted slope of ${\Delta r}/{\Delta E} = (+104\pm 25\mbox{(stat.+syst.)})\cdot 10^{-4}\ \mbox{GeV}^{-1}$. {\small \begin{figure}[hbt] \begin{center} \mbox{\epsfig{file=mont_mult.eps,height=11.cm,width=7.5cm}} \end{center} \caption{ (a) Mean charged multiplicity of quark and gluon jets and (b) multiplicity ratio $r$ as a function of the jet energy} \label{fig:proc_mult} \end{figure} } In a recent review \cite{bruseles} all published data from various experiments \cite{delphi_qg,opal_qg,aleph_qg,hrs_qg,sld_qg} were used to perform a general study of $r$ as a function of the jet energy. At present, more data can be added to this comparison. These are the new analysed DELPHI data sample presented above and the most recent measurements of $r$ performed by CLEO \cite{cleo_qg} and OPAL \cite{opal_hem} at 4-7 GeV and 39 GeV average jet energies, respectively. The updated new DELPHI analysis incorporates two times more statistics than the previous analysis \cite{delphi_qg}, therefore, significantly reduces the statistical errors. The analysis from CLEO compares the charged particle multiplicity in $\Upsilon(1S) \rightarrow gg\gamma$ decays to that observed in $e^+e^- \rightarrow q\overline{q}\gamma$ just in the continuum. This study does not rely on the Monte Carlo simulation to associate the final hadrons to the initial partons and can consequently be fairly considered as being model independent. The obtained value is $r=1.04\pm0.05\ \mbox{(stat.+syst.)}$. The OPAL analysis uses a new technique \cite{gary} which selects gluon jets at $\sim$39 GeV by dividing the events into two hemispheres. While one of these hemispheres is required to contain two tagged quark jets, the other is left untouched being regarded as the gluon jet. The result from OPAL, expressed for only light $uds$-quarks, is $r_{uds}=1.55\pm0.07\ \mbox{(stat.+syst.)}$. As it can be observed in figure \ref{fig:proc_mult}.b all these data agree with the predicted energy behaviour of \cite{delphi_qg,fodor,bruseles} when the correction to the quark multiplicity to account for the same flavour composition is applied. In our case it is: 56\% $uds$'s, 33\% $c$'s and 11\% $b$'s. The OPAL number considering this quark mixture becomes $r=1.48\pm0.07\ \mbox{(stat.+syst.)}$. All these results thus give evidence for an energy dependence of $r$. The measured increase is \[ \frac{\Delta r}{\Delta E} = (+110 \pm 13\ \mbox{(stat.+syst.)})\cdot10^{-4}\ \mbox{GeV}^{-1}, \] representing a $\sim$8$\sigma$ effect. The measured value of $r$ remains systematically lower than the {\sc Jetset} prediction over the whole energy range, having an average value of \[ r= 1.23\pm0.01\ (stat.) \pm0.03\ (syst.), \] which corresponds to an average energy of $\sim$23 GeV. This ratio can be further expressed as \[ r_{uds}= 1.30 \pm0.01\ (stat.) \pm0.04\ (syst.), \] if $r$ is computed only for the light $uds$-quarks, extracting the $b$ and $c$ quark contribution to the quark jet multiplicity. The absolute value of $r$ depends on the reconstruction jet algorithm. For both the {\sc Jade} and {\sc Cone} schemes different results are obtained w.r.t. the {\sc Durham} scheme \cite{delphi_qg,opal_qg}. This is due to the combined effect of the different sensitivity of the various jet reconstruction algorithms to soft particles at large angles and of the expected different angular and energy spectra of the emitted soft gluons in the quark and gluon jets. A precise deconvolution of both effects is, at present, impossible \cite{ruso}. This jet algorithm dependence of $r$ becomes however less apparent as the jet energy increases. The results from OPAL \cite{opal_hem}, $r=1.48\pm0.07\ \mbox{(stat.+syst.)}$ and those from DELPHI \cite{delphi_qg} at $\sim$40 GeV presented in this conference, $r=1.43\pm0.07\ \mbox{(stat.+syst.)}$ for {\sc Durham} and $r=1.52\pm0.11\ \mbox{(stat.+syst.)}$ for {\sc Jade}, agree within errors for the various methods and algorithms used. For the low energy interval, the {\sc Jade} and {\sc Durham} jet algorithms give a different description of the gluon jet properties \cite{delphi_qg}, although the {\sc Durham} algorithm is in better agreement to those, $model$ $independent$, results obtained by CLEO. Hence, the {\sc Durham} jet algorithm seems to be better suited to decribe the intermediate energy region than the {\sc Jade} algorithm is. The interpretation of these results in combination with those obtained by OPAL \cite{opal_bg} and ALEPH \cite{aleph_bg} restrict the validity of the statement that gluon and $b$-quark jets have similar properties to the jet energy interval around 24 GeV and cannot be applied to the whole jet energy spectrum. \section{GLUON RADIATION IN $b$-QUARKS} For many observable quantities at LEP energies, $\sqrt{s} \raisebox{0.5ex}{$>$}\hspace{-1.7ex}\raisebox{-0.5ex}{$\sim$}$M$_Z$, quark mass effects usually appear in terms proportional to m$_q^2/$M$_Z^2$. This represents a $\sim$3\raisebox{0.5ex}{\tiny $0$}$\!/$\raisebox{-0.3ex}{\tiny $\! 00$}{\normalsize \hspace{1ex}} correction for m$_q$=m$_b$ which in most of the cases can be savely neglected. This argument, for instance is true for the total hadronic cross section \cite{val} but cannot be applied for the differential multi-jet cross sections that depend on the jet-resolution parameter, $y_c$. The reason being the new scale, $E_c = M_Z \sqrt{y_c}$, introduced in the analysis by the new variable which enhances the mass effects in the form $m_b^2/E_c^2 = (m_b^2/M_Z^2)/y_c$. At $\sqrt{s} \approx$M$_Z$ the three-jet production rate for $b$-quarks is in fact suppressed by a factor $\sim5-10$\% w.r.t that of light quarks \cite{tor,val,delphi_ab}. This difference can then be expressed as a function of m$_b$ \cite{val} and, therefore, used to measure its value. \begin{figure}[ht] \begin{center} \mbox{\epsfig{file=systematics.eps,width=7.5cm} } \end{center} \caption{Relative systematics uncertainties in $R_3^{bd}$ due to fragmentation} \label{fig:systematics} \end{figure} The experimental observation of such effects is however difficult and delicate because the effect is after all small and furthermore the correct theoretical framework to resolve the mass definition ambiguities is needed. This means that the observable has to be calculated including mass effects at O($\alpha_s^2$). For this purpose a recent calculation \cite{german} of the ratio of the normalized three-jet cross sections between $b$-quarks and light $uds$-quarks \[ R_3^{bd} \equiv \frac{\Gamma_{3j}^{Z^0\rightarrow b\bar{b}g}(y_c)/ \Gamma_{tot}^{Z^0\rightarrow b\bar{b}}} {\Gamma_{3j}^{Z^0\rightarrow d\bar{d}g}(y_c)/ \Gamma_{tot}^{Z^0\rightarrow d\bar{d}}} \label{eq:r3bd} \label{rtheta} \] has been performed \begin{figure}[ht] \begin{center} \mbox{\epsfig{file=restes.eps,width=7.5cm} } \end{center} \caption{$R_3^{bd}$ distribution} \label{fig:r3bd} \end{figure} The normalization in $R_3^{bd}$ to the total decay rates is introduced to cancel possible weak corrections depending on the top quark mass \cite{top} and the ratio of the three-jet cross sections between $b$ and light $uds$-quarks minimizes uncertainties due to the hadronization process. In figure \ref{fig:systematics} the dependence of these uncertainties w.r.t the $y_{cut}$ is shown and seen not to exceed 3\raisebox{0.5ex}{\tiny $0$}$\!/$\raisebox{-0.3ex}{\tiny $\! 00$}{\normalsize \hspace{1ex}} for large enough values of $y_{cut}$. The $R_3^{bd}$ distribution corrected for detector and fragmentation effects is also displayed in figure \ref{fig:r3bd}. The solid curves drawn in the figure are the theoretical O($\alpha_s$) prediction in steps of 1 GeV. The values of m$_b$ used to produce these curves are meaningless since they correspond to a calculation at O($\alpha_s$). They can nevertheless be used to evaluate the experimental precision assuming the difference between the theoretical curves remains similar to that at O($\alpha_s^2$). As can be observed the experimental error corresponds then to approximately 300 MeV for reasonably high values of $y_{cut}$.
proofpile-arXiv_065-470
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} Topological field theories offer an intriguing possibility to combine ideas from physics and mathematics. They are quantum field theories with no physical degrees of freedom and their properties are fully determined by the global structure of the manifold they are defined on. A remarkable feature is that for many topological theories, like the Donaldson theory \cite{Wit-TQFT} and Chern-Simons (CS) theory, the expectation values of the observables are topological invariants. The Chern-Simons theory provides a three-dimensional interpretation of the theory of knots: the correlators of its observables, Wilson loops, are related to the Jones polynomials of knot theory \cite{Wit-Jones}. Another important application of CS theory is $2+1$ dimensional gravity. CS action with Poincar\'e group as the gauge group is the Einstein-Hilbert action \cite{2+1}, giving a gauge theory interpretation of gravity in $2+1$ dimensions. However, the Chern-Simons theory is defined only in three dimensions. Its generalization to arbitrary dimensions \cite{KaRo-arb,Horo-ex} are called BF-models or antisymmetric tensor models. They, like the CS theory, describe the moduli space of flat connections and their observables are related to the linking and intersection numbers of manifolds. The supersymmetric BF-theories (SBF) were introduced in \cite{Wit-topgra} as a supersymmetric version of $2+1$ dimensional topological gravity. There it was also shown that the partition function of three dimensional SBF computes a topological invariant, the Casson invariant. Generalizations of SBF to other dimensions were considered in \cite{BBT} and \cite{Wal-alg,Horo-ex}. In this article we study supersymmetric BF models. We are particularly interested in finding new observables and possible topological invariants for $3d$ SBF-theories, besides the partition function. By formulating the theory in superspace a large set of observables, including previously unknown ones, can be derived form the superspace curvature. In \cite{BiRa-JGeom} a vector-like supersymmetry similar to that found in ordinary BF-models and Chern-Simons theory \cite{DGS-3d,BRT-ren} was constructed for the SBF-models. In particular, the hierarchy of observables constructed from the supercurvature can be derived from one initial observable with the help of the vector supersymmetry. Using the superspace formulation we extend this construction to include also the anti-BRST and anti-vector supersymmetries, in addition to the usual BRST and vector supersymmetries. This article is organized as follows: in section 2 we will introduce the model and write it in the superspace. It turns out that in the superspace formalism many features of the CS theory can be generalized directly to SBF-theory. In section 3 we derive the set of observables and discuss their relation to topological invariants. In section 4 we generalize the vector supersymmetry to SBF and show how it can be used to construct of new observables. \section{Supersymmetric BF-theories} \la{SBF} The classical action or non-supersymmetric BF-models in $d$ dimensions is \begin{equation} S_0 = \int \d^dx\, B_n^0 F_A, \nonumber \end{equation} where $B_n^0$ is a $n=d-2$ form (with ghost number zero) and $F_A$ is the curvature two form $F_A = \d A + \frac{1}{2} [A,A]$. In addition to the normal Yang-Mills gauge symmetry $ A \to A + \d_A \omega_0$ the action is invariant under the transformation $ B_n \to B_n+ \d_A\omega_{n-1}$ caused by the Bianchi identity. In dimensions higher than three this symmetry is reducible: $$ \omega_{n-1} \to \d_A\omega_{n-2} \quad\hbox{\rm etc. } $$ and additional ghost fields are needed in order to fix the gauge according to the Batalin-Vilkovisky procedure. In three dimensions the BF theory is closely related to the Chern-Simons theory: CS-theory for the tangent group $TG \simeq (G, \underline{g})$ is equivalent to the BF-theory for $G$ \cite{BBRT-TQFT}. In $TG$ the Chern-Simons connection one-form splits into two parts $A$ and $B$, the basic fields of the BF-theory. This makes it possible to construct the classical action of BF-theories, find the BRST-transformations and fix the gauge easily by studying the CS theory for the tangent group. For the supersymmetric extension of three-dimensional BF-model the situation is quite similar --- the action and many properties of the theory can be expressed in terms of super CS-theory. This is done elegantly by formulating the theory in superspace with two anticommuting Grassmannian coordinates $\theta$, $\bar\theta$ in addition to the normal space time coordinates $x_\mu$. Here we will mainly concentrate in the three-dimensional case but with slight modifications the method is suited for SBF-models in other dimensions. The integration over the Grassmannian variables is normalized as \begin{equation} \int \d\bar\theta\d\theta \left\{\begin{array}{c} 1 \\ \theta \\ \bar\theta \\ \theta\bar\theta \end{array}\right\} = \left\{\begin{array}{c} 0\\ 0\\ 0\\ 1 \end{array}\right\}. \end{equation} If the coordinates $\theta$ and $\bar \theta$are associated with ghost numbers $-1$ and 1 the superspace connection one-form $\hat{\cal A}$ in $3+2$ dimensions is written\footnote{ Note that we will use graded differential forms $X^q_p$ with ordinary form degree $p$ and ghost number $q$. Two graded forms satisfy $X^q_p Y^r_s = (-1)^{(q+p)(r+s)}Y^r_sX^q_p $. All the commutators should also be considered graded.} \begin{equation} \hat{{\cal A}} = \hat{ A}^0_\mu\d x^\mu + \hat{A}^1_\theta \d\theta + \hat{A}^{-1}_{\bar\theta}\d\bar\theta \la{suconn} \end{equation} where the superfields $\hat A^0_\mu\d x^\mu ,\ \hat A^1_\theta $ and $\hat A^{-1}_{\bar\theta}$ can be further expanded as: \begin{eqnarray} \hat A^0_\mu &=& A_\mu -\theta\psi_\mu + \bar\theta\chi_\mu + \theta\bar\theta B_\mu\nonumber\\ \hat A^1_\theta &=& c -\theta\phi + \bar\theta\rho + \theta\bar\theta\eta\nonumber\\ \hat A^{-1}_{\bar\theta} &=& \bar\eta -\theta\bar\rho - \bar\theta\bar\phi + \theta\bar\theta \bar c.\nonumber \end{eqnarray} The components can be identified with the fields of three dimensional super BF-theory: $\psi^1_\mu$ and $\chi^{-1}_\mu$ are the superpartners of the connection $A^0_\mu$ and field $B^0_\mu$, while $ \rho^0_0, \bar\rho^0_0$ and $\phi^2_0, \bar\phi^{-2}_0$ are their corresponding ghosts and antighosts. With these definitions the classical action of the SBF-model can be written as the action of the super-CS theory: \begin{equation} S_{cl} = \int \d^3x (BF_A - \chi\d_A\psi)= \frac{1}{2}\int\d^3x \d^2\theta (\hat{\cal A} \hat\d\hat{\cal A} + \frac{2}{3} \hat{\cal A}[\hat{\cal A},\hat{\cal A}] ) . \la{SBFcl} \end{equation} To obtain the quantum action one has to fix the gauge symmetry $\hat A^0 \to \hat A^0 + \d_{\hat A^0} \omega$ by adding to the action a BRST-exact gauge fixing term. The BRST transformations of the fields can be derived from the superspace curvature two-form using a method similar to that of \cite{BiRa-JGeom,Wal-alg,MaiNie,HiNiTi,unicon,BiRaTh-red} for Donaldson theory and Witten type topological theories. But because of the $N=2$ superspace with two anticommuting coordinates of opposite ghost numbers we can extend this method to include also the anti-BRST symmetry. We define the superspace curvature as \begin{equation} \hat{\cal F} = (\d x^\mu \partial_\mu + \d\theta \delta +\d\bar\theta \bar\delta ) \hat {\cal A} + \frac{1}{2} [\hat{\cal A}, \hat {\cal A} ] \la{scurv} \end{equation} and impose the `horizontality condition'' \begin{equation} \hat{\cal F} \equiv \hat F_{\mu\nu}\d x^\mu \d x^\nu - (\d\theta\partial_\theta + \d\bar\theta\partial_{\bar\theta})\hat {\cal A} , \la{restrict} \end{equation} which truncates the curvature to the physical part independent of $\d\theta,\ \d\bar\theta$ (and consequently of the ghost fields), and identifies the BRST-operator $\delta$ with $\partial_\theta$ and $\bar\delta$ with $\partial_{\bar\theta}$. This gives the BRST-transformations for the component fields: \begin{equation} \begin{array}{llll} \delta A &=-\d_A c +\psi \quad &\delta B &= -\d_A \eta - [c,B] + [\phi,\chi]+ [\psi,\rho]\cr \delta c &= -\frac{1}{2} [c,c] +\phi \quad &\delta \eta &=-[c,\eta] +[\phi,\rho] \cr \delta \psi &= -\d_A \phi - [c,\psi] \quad &\delta \chi &= -\d_A\rho - [c,\chi] +B\cr \delta \phi &= -[c,\phi] \quad &\delta \rho &= -[c,\rho] +\eta \end{array} \la{BRST1} \end{equation} which have to be supplemented with the transformations of the anti-ghosts and Lagrange multipliers $\lambda_0^0, b_0^0, \beta_0^{-1}, \sigma_0^1$ for the gauge fixing conditions of the fields $A,B,\psi$ and $\chi$. The Lagrange multipliers can be combined into a superfield \begin{equation} \hat\Lambda_0^0 = \lambda -\theta \sigma -\bar\theta\beta +\theta\bar\theta b . \la{lambda} \end{equation} The simplest choice for the BRST transformations would be $$ \delta \hat{\cal A}^{-1}_0 = -\hat\Lambda, \qquad\delta \hat\Lambda=0 $$ but with suitable field redefinitions these can be put into a form which will be more convenient later on: \begin{equation} \begin{array}{llll} \delta \bar c &= -b \quad &\delta \bar\eta&= -\lambda -[c,\bar\eta] + \bar\rho\\ \delta b &= 0\quad &\delta\lambda &= -[c,\lambda] -[\phi,\bar\eta] -\sigma \\ \delta \bar\phi &= \beta - \bar c \quad &\delta \bar\rho&= \sigma -[c,\bar\rho] \\ \delta \beta &= -b \quad &\delta \sigma &= -[c,\sigma] +[\phi,\bar\rho]. \end{array} \la{BRST2} \end{equation} The gauge fixing part of the supersymmetric action is chosen to be \begin{eqnarray} S_{gf} &=& \int \d^2\theta\, \delta(\hat A^{-1}_{\bar\theta} \d * \hat A^0 )\cr &=&\int \d^3x \,(-b \d *A -\lambda \d *B + \beta\d *\psi + \sigma\d *\chi + \bar c \d *\d_A c +\bar \eta \d *\d_A \eta \la{SBFgf}\\ &+& \bar \phi\d *\d_A \phi + \bar \rho \d *\d_A \rho - \bar\eta [\d c, *B] + \bar\eta \d [\phi, *\chi] - \bar\eta \d [\rho, *\psi]+ \bar\rho [\d c, *\chi] +\bar\phi \d [c, *\psi]) \nonumber \end{eqnarray} Note also that unlike in the ordinary BF-model the classical action \nr{SBFcl} is now BRST-exact \[ S_{cl} = \int \d^3x\, \delta(\chi F_A ). \] This shows that the supersymmetric BF model is a Witten type topological theory with a $\delta$-exact action, whereas the ordinary non-abelian BF models are Schwartz type theories \cite{BBRT-TQFT}. The gauge fixing term $S_{gf}= \int \d^2\theta\, \delta(\hat A^{-1}_{\bar\theta} \d * \hat A^0)$ is formally similar to that of Chern-Simons theory quantized in Landau gauge $\d *{\cal A} =0$: \[ S^{CS}_{gf}= \int \d^3 x (\delta \bar {\cal C} \d * {\cal A}) \] In CS theory the BRST and anti-BRST operators are related by transformation obtained by integrating the quantum action by parts \cite{BiRa-vecSUSY} \[ S^{CS}_q = \int \d^3 x ({\cal A} \d{\cal A} + \frac{2}{3} {\cal A}[{\cal A},{\cal A}] - \Lambda \d * {\cal A} - \bar {\cal C} \d *\d_{\cal A} {\cal C}). \] The integrated action is equivalent to the original action after a change of fields which leaves the connection ${\cal A}$ unchanged but takes the ghosts ${\cal C}$ to the antighosts $\bar{\cal C}$ and $\bar{\cal C}$ to $-{\cal C}$. The Lagrange multiplier field $\Lambda$ transforms as $\Lambda \to \Lambda - [{\cal C}, \bar{\cal C}]$. This transformation of the fields maps $\delta$ to $\bar\delta$ . For the super-CS and consequently for the three dimensional SBF-theory the situation is again analogous. Integrating the gauge fixed quantum action $S_{q} = S_{cl}+ S_{gf}$ (\ref{SBFcl},\ref{SBFgf}) by parts we find the superspace version of the transformation which relates BRST- and anti-BRST operators. In superspace language the transformation rules can be expressed compactly by demanding that under the ``conjugation'' of the Grassmann variables \begin{equation} \theta \to \bar \theta,\qquad \bar\theta \to - \theta \la{conjugation} \end{equation} the total superspace connection $\hat{\cal A}$ stays the same while the operators change as $\delta\to\bar\delta,\ \bar\delta\to -\delta$. The transformations for the Lagrange multipliers are somewhat more complicated \begin{equation} \begin{array}{llll} \lambda &\to \lambda + [c, \bar\eta]\qquad & b &\to b -[c, \bar c]-[\eta , \bar \eta ] -[\rho , \bar\rho ]-[\phi ,\bar\phi] \cr \sigma &\to \beta + [c, \bar\phi] \qquad &\beta &\to -\sigma +[\phi,\bar\eta]. \la{trans} \end{array} \nonumber \end{equation} For BF-theories in dimensions other than three the situation is slightly more complicated because the $A$ and $B$ fields cannot be combined into one connection. In $d$ dimensions $B$ is a $d-2$ form and additional fields will be needed to take care of the reducibility. It is however possible to use truncated fields and write the components of $\hat{\cal A}$ as \begin{eqnarray} \hat A_\mu^0 &=& A_\mu - \theta \psi_\mu \cr \hat A_\theta^1 &=& c - \theta \phi \cr \hat A_{\bar\theta}^{-1} &=& -\bar\theta\bar\phi + \theta\bar\theta \bar c \la{trunc} \end{eqnarray} and similarly for the $d-2$ superform $\hat{\cal B}$, which now contains in addition to $B$, $\chi$ and their ghosts also the whole tower of ghosts for ghosts from the Batalin-Vilkovisky gauge fixing. The curvature of the $B$ sector is defined as \begin{equation} \hat{\cal R} =(\d x^\mu \partial_\mu + \d\theta\delta )\hat{\cal B} + [\hat{\cal A} , \hat{\cal B} ]. \end{equation} It satisfies a Bianchi identity, and again after imposing the horizontality condition similar to \nr{restrict} it reproduces the correct nilpotent BRST-transformations. Since the $A$ and $B$ sectors do not appear symmetrically in the action there exists no partial integration symmetry and thus no anti-BRST operator $\bar\delta$. \section{Observables} \la{Obs} In order to establish that the observables of the theory are indeed topological invariants it must be checked that they are BRST-closed, their expectation values do not depend on variations of the metric and, if they are integrals of some local functionals, that their BRST-cohomology depends only on homology class of the integration contour. The partition function of the three dimensional SBF-model $$ Z_{3d} = \int e^{iS_q}. $$ obviously satisfies all the requirements and can be shown to equal the Casson invariant of the manifold \cite{Wit-topgra,BlaTho-Casson}. We will now derive a set of other observables for $3d$ SBF form the superspace curvature \nr{scurv} and see if they too could correspond to topological invariants. The Bianchi identity \begin{equation} (\d x^\mu \partial_\mu +\d\theta\delta + \d\bar\theta\bar\delta)\hat{\cal F} + [\hat{\cal A},\hat{\cal F}] =0 \nonumber \end{equation} guarantees that the powers of $\hat{\cal F}$ obey \begin{equation} (\d x^\mu \partial_\mu +\d\theta\delta + \d\bar\theta\bar\delta ) \hbox{\rm Tr\ } {\hat {\cal F}}^n =0. \la{trcurv} \end{equation} The simplest one is the superspace 4-form ${\hat{\cal F}}^2$. It can be expanded in powers of $\d\theta$ and $\d\bar\theta$: \begin{equation} \frac{1}{2} \hbox{\rm Tr\ } {\hat{\cal F}}^2 = \sum_{i,j; i+j\le 4} W^{i,j} {\d\theta}^i{\d\bar\theta}^j . \nonumber \end{equation} Equation \nr{trcurv} gives \begin{equation} \d W^{i,j} + \delta W^{i-1,j} + \bar\delta W^{i,j-1} = 0. \la{cohom} \end{equation} When $j=0$ the integrals of the $(4-i)$-form $W^{i,0}$ over a $(4-i)$ cocycle $\gamma$ are BRST-closed: \begin{equation} \int_\gamma \d W^{i,0} + \delta \int_\gamma W^{i-1,0} = \int_{\partial\gamma} W^{i,0} + \delta \int_\gamma W^{i-1,0} = \delta\int_\gamma W^{i-1,0} =0. \nonumber \end{equation} Because of \nr{cohom} the BRST-cohomology of $\int W$ depends only on the homology class of $\gamma$, making the vacuum expectation values and correlation functions of $\int W$ good candidates for topological invariants. Note that because of the symmetry of the three-dimensional action (\ref{SBFcl},\ref{SBFgf}) under the partial integration transformation the expectation values of any observable $\O $ and its ``conjugate'' $\overline{\O}$ are the same. This can be seen by making a change of variables (with a unit Jacobian) in the path integral taking all the fields to their conjugates and using the invariance of the action. The condition $\delta \O =0$ changes under this transformation to $\bar\delta \overline{ \O} =0$. Therefore objects that are either $\delta$- or $\bar\delta$-closed qualify as observables of $3d$ SBF. In particular, we can thus identify $\overline{W}^{i,j} = (-1)^j W^{j,i}$. The expansion of ${\hat{\cal F}}^2$ gives using \nr{restrict} \begin{eqnarray} W^{00} &=& \frac{1}{2} F^2 \cr W^{10} &=& \psi F - \bar\theta (BF - \chi \d_A \psi)\cr W^{20} &=& \frac{1}{2} \psi^2 +\phi F + \theta (\phi\d_A \psi) - \bar\theta (\psi B + \phi\d_A\chi + F\eta)\cr &+& \theta\bar\theta ( \phi \d_A B - \phi [\psi,\chi] + \d_A \psi \eta) \cr W^{30} &=& \psi\phi -\bar\theta (\phi B + \psi\eta) \cr W^{40} &=& \frac{1}{2} \phi^2 -\bar\theta (\phi\eta) \la{W} \end{eqnarray} {}from which we can extract 11 observables. The previously unknown ones are the $\theta$- and $\theta\bar\theta$-components of $W^{20}$. They are a particular to three-dimensional theories and unlike the others which, or rather their generalizations involving all the Batalin-Vilkovisky ghosts, can be obtained from the truncated supercurvatures $\hat{\cal F}$ and $ \hat{\cal R}$ of $A$ and $B$-sectors in all dimensions. Nevertheless, the $\theta$ component of $W^{20}$ seems to be BRST-closed also in higher dimensions: the ghosts for ghosts and other fields appear only in the transformations for the $B$-sector. Interestingly, some of the observables above are formally the same as for Donaldson theory. This is no surprise since the BRST-structure of Donaldson theory is similar to that of the $A$-sector of the SBF. In fact, the SBF can be thought as reduction of the Donaldson theory to three dimensions \cite{BiRaTh-red,MaiNie}. As characteristic for the Witten type topological theories the expectation values of observables \begin{equation} <\O> = \int [dX] \O e^{i/g^2 S_q} \nonumber \end{equation} are independent of the coupling $g^2$. The integral can be calculated in the $g^2 \to 0$ limit where it localizes to the classical equations of motion \begin{equation} F_A=0, \qquad \d_A \psi =0, \qquad \d_A B - [\psi,\chi]=0,\qquad \d_A\chi =0 \la{EQM} \end{equation} {\it i.e.\, }it is now calculated over the moduli space of flat connections ${\cal M}$. In the limit $g^2 \to 0$ the fields in \nr{W} are replaced by their classical values \nr{EQM}. Then the non-vanishing observables are \begin{equation} \begin{array}{llllll} \omega_0^4 &= \frac{1}{2}\phi^2\quad &\omega_1^3 &= \int\psi\phi \quad &\omega_2^2 &= \frac{1}{2}\int \psi^2\\ \omega_0^3 &= \phi\eta \quad &\omega_1^2 &=\int \psi\eta + B\phi\quad &\omega_2^1 &= \int\psi B. \end{array} \nonumber \end{equation} To evaluate the expectation values one has to take care of the zero modes of the fermions. Especially in dimensions higher than two the zero-modes of the other fields complicate matters considerably. We will not perform the calculations here but refer to \cite{BBRT-TQFT} and references therein for discussions on topological invariants of the Donaldson theory. The considerations there are quite similar to those for the observables $\omega_0^4,\ \omega_1^3,\ \omega_2^2$ of the $A$-sector of SBF. The invariant corresponding to $\omega^2_2$ has been evaluated in \cite{BBRT-TQFT} for $2d$ BF and found to be the symplectic volume of the moduli space. Its products with $\omega^4_0$ produce linking and intersection numbers of moduli spaces. \section{Vector supersymmetry and the tower of observables} \la{VecSUSY} A peculiar feature of Chern-Simons and BF-theories is a vector-like supersymmetry of the action \cite{BRT-ren}. It gives rise to new Ward identities which have been utilized in proving the theories to be finite, renormalizable and free of anomalies \cite{vecSUSY-1, vecSUSY-2}. However, this supersymmetry is valid only in flat space. One might argue though that because the theories are topological their physical quantities are not dependent on the metric of the manifold. In any case, the vector supersymmetry has been established as a common feature of many topological theories \cite{BiRa-JGeom} and a useful tool, not only in the study of renormalization and related topics but also in finding new observables. The vector supersymmetry for non-supersymmetric CS theory quantized in Landau gauge is generated by operator $s$ with ghost number and form degree 1: \begin{equation} \begin{array}{llll} s {\cal A} &= *\d{\cal C} \qquad &s {\cal C} &= 0 \nonumber\\ s \bar {\cal C} &= {\cal A} \qquad &s \Lambda &= -\delta {\cal A}, \end{array} \la{vectorSUSY} \end{equation} or written in component form \[ s = s_\alpha\d x^\alpha \qquad *\d {\cal C} =-\epsilon_{\mu\alpha\beta} \partial^\beta{\cal C} \d x^\mu\d x^\alpha. \] Using the partial integration for the CS theory one can obtain the anti-supersymmetry $\bar s$: \begin{equation} \begin{array}{llll} \bar s {\cal A} &= *\d\bar{\cal C} \qquad & \bar s \bar{\cal C} &= 0 \nonumber\\ \bar s {\cal C} &= -{\cal A} \qquad & \bar s \Lambda &= - \bar\delta {\cal A} - [{\cal A},\bar {\cal C}] \end{array} \la{avectorSUSY} \end{equation} The anti\-commutation relations of the operators $\delta,\bar\delta$ and $s, \bar s$ are \begin{eqnarray} [s_\alpha , s_\beta ]&=&[\bar s_\alpha , \bar s_\beta ]= [\delta, \bar\delta ] = [s_\alpha ,\bar s_\beta ] = [\delta , s_\alpha ] = [\bar\delta, \bar s_\alpha ] =0 \la{commu} \\ {} [\delta, \bar s_\alpha ] &=& -[\bar\delta , s_\alpha ] = \partial_\alpha + \hbox{\rm \, terms vanishing modulo the equations of motion}. \nonumber \end{eqnarray} Together with the BRST-operators $\delta$ and $\bar\delta$ the operators $s$ and $\bar s$ can be combined to form a generator of $N=2$ supersymmetry algebra \cite{DGS-3d,BiRa-vecSUSY,BiRa-JGeom}. The vector supersymmetries can be formulated also for the non-supersymmetric BF-theories. In dimensions other than three there exists no vector supersymmetry $s$ but the $\bar s$ operator can still be constructed \cite{vecSUSY-1}. The anti-vector supersymmetry can be generalized to the supersymmetric BF-theory in arbitrary dimensions. In $3d$ it can be derived easily using \nr{avectorSUSY} for the superfields $\hat A^0, \hat A^1_\theta, \hat A^{-1}_{\bar\theta}, \hat\Lambda$: \begin{equation} \begin{array}{llllll} \bar s_\alpha A_\mu &=&-\epsilon_{\mu\alpha\beta}\partial^\beta\bar\eta\quad &\bar s_\alpha B &=& -\epsilon_{\mu\alpha\beta}\partial^\beta\bar c \cr \bar s_\alpha c &=& A_\alpha\quad &\bar s_\alpha \eta &=& B_\alpha\cr \bar s_\alpha \bar c &=&0 \quad &\bar s_\alpha \bar\eta &=&0 \cr \bar s_\alpha b &=& -\partial_\alpha \bar c\quad &\bar s_\alpha \lambda &=& D_\alpha \bar\eta\cr \bar s_\alpha \psi_\mu &=& \epsilon_{\mu\alpha\beta}\partial^\beta\bar\rho \quad &\bar s_\alpha \chi &=&\epsilon_{\mu\alpha\beta}\partial^\beta\bar\phi \cr \bar s_\alpha \phi &=& -\psi_\alpha\quad &\bar s_\alpha \rho &=& -\chi_\alpha \cr \bar s_\alpha \bar \phi &=&0 \quad &\bar s_\alpha \bar\rho &=&0 \cr \bar s_\alpha \beta &=& \partial_\alpha \bar \phi\quad &\bar s_\alpha \sigma &=& D_\alpha \bar \rho \end{array} \la{aSUSY} \end{equation} This is a symmetry of the quantum action (\ref{SBFcl},\ref{SBFgf}) and satisfies the anticommutation relations \nr{commu} with the BRST-operator (\ref{BRST1},\ref{BRST2}). The analysis done on the renormalization, finiteness and anomalies of ordinary BF-theories using vector supersymmetry can thus be applied directly to the supersymmetric BF-theories. It is interesting to note that the vector supersymmetry of the SBF can be useful also in constructing new observables (see also \cite{BiRa-JGeom} for a slightly different approach). Whenever there exists a BRST-closed object $\omega$ also $\bar s\omega$ is BRST-closed as a result of the anticommutation relations \nr{commu}. So in principle it is possible to find an observable, like $\omega_0^4$ and $\omega_0^3$, and apply $\bar s_\alpha$ successively to get new ones. Also, since \begin{equation} \omega_0^4 = \frac{1}{2}\delta (c\phi -\frac{1}{6}c[c,c]),\qquad \omega_0^3 = \frac{1}{2}\delta ( \phi \rho +c\eta -\frac{1}{2} \rho [c,c] )= \delta (\phi\rho) \nonumber \end{equation} all observables obtained by acting with $\bar s$ are in fact BRST-exact --- modulo equations of motion and surface terms. This is valid only locally and does not mean that the observables should be trivial. {}From \nr{W} we see that modifying slightly the anti-supersymmetry transformations for $\psi$ and $B$ in \nr{aSUSY} as \[ \bar s B = *\d \bar c -2\d_A \chi, \qquad \bar s\psi = -*\d\bar\rho -2 F \] and leaving the others intact the anti-supersymmetry still remains a symmetry of the action. By denoting the metric independent part of the modified $\bar s$ operator by $\bar v$: \begin{equation} \begin{array}{llllll} \bar v B &= -2\d_A\chi \quad &\bar v\eta &= -B\quad &\bar v\rho &= -\chi \cr \bar v\psi&= -2F \quad &\bar v c &= -A\quad &\bar v\phi &= -\psi \end{array} \la{mods} \end{equation} and applying successively $\frac{1}{k!}(-\bar v)^k$ to $\frac{1}{2}\phi^2$ and $\phi\eta$ it is possible to derive all the observables in \nr{W}, except the $\theta$ and $\theta\bar\theta$ components of $W^{02}$. The symmetry $\bar v$ acts as a vertical (in the direction of the form degree) transformation along the components $W^{ij}$ of the supersurvature \nr{scurv}. It is easily seen that also a horizontal (ghost number direction) transformation $\bar h$ can be defined: \begin{equation} \begin{array}{llll} \bar h A &=\chi \quad &\bar h c &= -\rho\cr \bar h \psi &= -B \quad &\bar h\phi &= -\eta\cr \bar h\bar\eta &=-\bar\phi\quad &\bar h\bar\rho &= -\bar c . \end{array} \la{horizontal} \end{equation} This is a symmetry of the action and commutes with the BRST-operator. It thus allows us to construct all possible observables starting from the element $\frac{1}{2} \phi^2$ of highest ghost number and lowest form degree --- again excluding the $\theta$ and $\theta\bar\theta$ components of $W^{20}$. The vertical transformation has geometrical interpretation as the equivariant derivative of BRST-model acting on the curvature of the universal bundle over the space of gauge connections \cite{BiRa-uni}, which can be identified with the supercurvature $\hat F$. Therefore the vertical transformation can be defined for all Witten type topological theories. The horizontal transformation $\bar h$ which can be constructed only in three dimensions is in fact part of the anti-BRST operator $-\delta$: only those terms that are not composites of fields and do not contain Lagrange multipliers are included. Acting on $\phi^2$ with the total vector supersymmetry transformation $\bar s$ instead of the vertical transformation we get even a larger set of observables. In addition to theose present in \nr{W} these include observables that depend explicitely on the metric. Since we already know that the observables in \nr{W} are BRST-closed also the metric dependent ones should be closed separately. Moreover, it can be shown that the metric variations of these observables can be written as BRST-exact terms, a necessary requirement for the observables to be topological invariants \cite{BBRT-TQFT}. So we can conclude that the expectation values of these observables are of topological nature. \section{Conclusions} We have studied three dimensional supersymmetric BF-theories using superspace formalism. This proved out to be a powerful method for studying the properties of the theory and especially for finding new symmetries and observables. The superspace curvature gives rise to a hierarchy of observables, which could be derived starting from one initial observable using the two transformations we constructed. The transformations have a geometrical interpretation as vertical and horizontal transformations acting on the components of the supercurvature, and can be identified as parts of more general symmetries of the action, the vector supersymmetry and anti-BRST symmetry. \subsection*{Acknowledgements} We thank prof. A. J. Niemi for useful comments on the manuscript and the referee for pointing out \cite{BiRa-JGeom}. \baselineskip 0.4cm
proofpile-arXiv_065-471
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} Some charge-transfer salts are understood as strongly correlated one-dimensional electron systems for half-filled bands (Mott-Hubbard insulators) (Farges 1994); (Alc\'{a}cer, Brau, and Farges 1994). The extended Hubbard model for interacting electrons on a Peierls-distorted chain at half-filling is considered appropriate for these materials (Mazumdar and Dixit 1986); (Fritsch and Ducasse 1991); (Mila 1995). There are only few studies of the optical, i.~e., finite frequency properties of correlated electron systems since their calculation is a formidable task (Kohn 1964); (Maldague 1977); (Lyo and Galinar 1977); (Lyo 1978); (Galinar 1979); (Campbell, Gammel, and Loh 1988); (Mahan 1990); (Shastry and Sutherland 1990); (Stafford, Millis, and Shastry 1991); (Fye, Martins, Scalapino, Wagner, and Hanke 1992); (Stafford and Millis 1993). In this second article on the optical absorption of electrons in half-filled Peierls-distorted chains we present a detailed analysis of the optical absorption in the limit of strong correlations and for a half-filled band where the charge and spin dynamics decouple. We find that the physics of the half-filled Hubbard model at strong correlations is determined by the upper and lower Hubbard band for the charges which are {\em parallel\/} bands. This essential feature and its significant consequences have been missed in earlier analytical and numerical investigations. In this work we include a finite lattice dimerization and nearest-neighbor interaction between the electrons, i.~e., we analyze the extended dimerized Hubbard model at strong correlations. The paper is organized as follows. In section~\ref{Hamilts} we address the Hubbard model at strong coupling from which we derive the Harris-Lange model which determines the motion of the charge degrees of freedom. In the ground state there are no free charges but only singly occupied lattice sites. The exact spectrum and eigenstates of the Harris-Lange model are presented in section~\ref{exactsolution} for the translational invariant and the dimerized case. The results can be interpreted in terms of two {\em parallel\/} Hubbard bands for the charges which are eventually split into Peierls subbands. Optical absorption can now formally be treated as if we had {\em independent\/} (spinless) Fermions. Unfortunately, this simple band structure interpretation is blurred by the spin degrees of freedom which enter the expressions for the optical absorption in terms of a very complicated ground state expectation value. In section~\ref{optabsHL} we treat the case of the Harris-Lange model where all spin configurations are equally possible ground states. This corresponds to the Hubbard model at temperatures large compared to the spin exchange energy. It allows the calculation of the optical absorption even in the presence of a Peierls distortion and a nearest-neighbor interaction between the charges. A summary and outlook closes our presentation. Some details of the calculations are left to the appendices. \section{Strongly correlated Mott-Hubbard insulators} \label{Hamilts} \subsection{Tight-binding electrons on a Peierls-distorted chain} \label{Peidistrot} For narrow-band materials the electron transfer is limited to nearest neighbors only. In standard notation of second quantization the Hamiltonian for electrons in the tight-binding approximation reads \begin{equation} \hat{T}(\delta)= - t \sum_{l=1,\sigma}^{L} \left(1+ (-1)^l\delta\right) \left( \hat{c}_{l,\sigma}^+ \hat{c}_{l+1,\sigma}^{\phantom{+}} + \hat{c}_{l+1,\sigma}^+ \hat{c}_{l,\sigma}^{\phantom{+}} \right) \end{equation} where $\delta$ describes the effect of bond-length alternation on the electron transfer amplitudes. As usual the Hamiltonian can be diagonalized in momentum space. We apply periodic boundary conditions, and introduce the Fourier transformed electron operators as $\hat{c}_{k,\sigma}^+=\sqrt{1/L} \sum_{l=1}^L \exp(ikla) \hat{c}_{l,\sigma}^+$ for the~$L$ momenta~$k=2\pi m/(La)$, $m=-(L/2),\ldots (L/2)-1$. We may thus write \begin{equation} \hat{T}(\delta) = \sum_{|k|\leq \pi/(2a),\sigma} \epsilon(k)\left( \hat{c}_{k,\sigma}^+\hat{c}_{k,\sigma}^{\phantom{+}} - \hat{c}_{k+\pi/a,\sigma}^+\hat{c}_{k+\pi/a,\sigma}^{\phantom{+}} \right) -i \Delta(k) \left( \hat{c}_{k+\pi/a,\sigma}^+\hat{c}_{k,\sigma}^{\phantom{+}} - \hat{c}_{k,\sigma}^+\hat{c}_{k+\pi/a,\sigma}^{\phantom{+}}\right) \label{Tink} \end{equation} with the dispersion relation~$\epsilon(k)$ and hybridization function~$\Delta(k)$ defined as \begin{mathletters} \begin{eqnarray} \epsilon(k) &=& -2t\cos(ka)\label{gl3a} \\[3pt] \Delta(k) &=& 2t\delta\sin(ka)\; . \end{eqnarray} \end{mathletters}% The Hamiltonian can easily be diagonalized in $k$-space. The result is (Gebhard, Bott, Scheidler, Thomas, and Koch I 1996) \begin{equation} \hat{T}(\delta) = \sum_{|k|\leq \pi/(2a),\sigma} E(k) (\hat{a}_{k,\sigma,+}^+ \hat{a}_{k,\sigma,+}^{\phantom{+}} - \hat{a}_{k,\sigma,-}^+ \hat{a}_{k,\sigma,-}^{\phantom{+}}) \; . \label{Tdiapeierls} \end{equation} Here, $\pm E(k)$ is the dispersion relation for the upper~($+$) and lower~($-$) Peierls band, \begin{equation} E(k) = \sqrt{\epsilon(k)^2 + \Delta(k)^2} \; . \label{peierlsen} \end{equation} The new Fermion quasi-particle operators $\hat{a}_{k,\sigma,\pm}^{+}$ for these two bands are related to the original electron operators by \begin{mathletters} \label{mixamp} \begin{eqnarray} \hat{a}_{k,\sigma,-}^{\phantom{+}} &=& \alpha_k \hat{c}_{k,\sigma}^{\phantom{+}} + i \beta_k \hat{c}_{k+\pi,\sigma}^{\phantom{+}} \\[6pt] \hat{a}_{k,\sigma,+}^{\phantom{+}} &=& \beta_k \hat{c}_{k,\sigma}^{\phantom{+}} -i \alpha_k \hat{c}_{k+\pi,\sigma}^{\phantom{+}} \end{eqnarray} \end{mathletters}% with \begin{mathletters} \label{alphabeta} \begin{eqnarray} \alpha_k &=& \sqrt{ \frac{1}{2} \left( 1 - \frac{\epsilon(k)}{E(k)} \right) }\\[6pt] \beta_k &=& - \sqrt{ \frac{1}{2} \left( 1 + \frac{\epsilon(k)}{E(k)} \right)} {\rm sgn}\left(\Delta(k)\right) \end{eqnarray} \end{mathletters}% Details of the upper transformation and the optical absorption of this model are presented in (Gebhard {\em et al.} I 1996). \subsection{Hubbard model} \label{Hubbard-Model} The only spinful interacting electron model that can be solved exactly for all values of the interaction strength is the Hubbard model in one dimension (Hubbard 1963); (Gebhard and Ruckenstein 1992); (E\ss ler and Korepin 1994); (Gebhard, Girndt, and Ruckenstein 1994); (Bares and Gebhard 1995). For narrow-band materials the electron transfer is limited to nearest neighbors only, and the interaction is supposed to be described by the purely local (Hub\-bard-)interaction of strength~$U$, \begin{eqnarray} \hat{H}_{\rm Hubbard}&=& \hat{T} +U \hat{D}\nonumber\\[6pt] \label{Hubb-Model} \hat{D}&=&\sum_{l} \hat{D}_{l} = \sum_l \hat{n}_{l,\uparrow}\hat{n}_{l,\downarrow} \; , \end{eqnarray} where $\hat{n}_{l,\sigma}=\hat{c}_{l,\sigma}^+\hat{c}_{l,\sigma}^{\phantom{+}}$ is the local density of $\sigma$-electrons, and $\hat{T}=\hat{T}(\delta=0)$. The model~(\ref{Hubb-Model}) poses a very difficult many-body problem. Its spectrum and, in particular, its elementary excitations can be obtained from the Bethe Ansatz solution (Lieb and Wu 1968); (Shastry, Jha, and Singh 1985); (Andrei 1995). Its low-energy properties including the DC-conductivity, $\sigma_{\rm DC}={\rm Re}\{\sigma(\omega=0)\}$, can explicitly be obtained from the corresponding $g$-ology Hamiltonian (Schulz 1990); (Schulz 1991) or from conformal field theory (Frahm and Korepin 1990); (Frahm and Korepin 1991); (Kawakami and Yang 1990); (Kawakami and Yang 1991). The Hubbard model describes a (correlated) metal for all~$U>0$ for less than half-filling. Unfortunately, the Bethe Ansatz solution does not allow the direct calculation of transport properties at finite frequencies. At half-filling the one-dimensional Hubbard model describes a Mott-insulator which implies that~$\sigma_{\rm DC}=0$ for all~$U>0$. The density of states for charge excitations displays two bands, the upper and lower Hubbard band, separated by the Mott-Hubbard gap. This gap is defined as the jump in the chemical potential at half filling, \begin{mathletters} \begin{eqnarray} \Delta_{\rm MH}&=&\mu^+(N=L)-\mu^-(N=L) \nonumber \\[6pt] &=& \left[ E(N=L+1)-E(N=L)\right] - \left[ E(N=L)-E(N=L-1)\right] \quad. \end{eqnarray} As shown by Ovchinnicov (Ovchinnicov 1969) the Mott-Hubbard gap can be obtained from the Lieb-Wu solution (Lieb {\em et al.} 1968) in the form \begin{eqnarray} \Delta_{\rm MH} &=& \frac{16t}{U} \int_{1}^{\infty} \frac{dy \sqrt{y^2-1}}{\sinh(2\pi t y/U)} \\[9pt] &=& \left\{ \begin{array}{ccr} (2W/\pi) \sqrt{4U/W} \exp(-\pi W/(2U)) & \hbox{for} & U \ll W=4t \\[6pt] U - W + \ln(2)W^2/(2U) + {\cal O}(W^3/U^2) & \hbox{for} & U \gg W=4t \end{array} \right. \quad . \end{eqnarray} \end{mathletters}% It is obvious that optical absorption is only possible if $\omega \geq \Delta^{\rm MH}$. It is further seen that the upper and lower Hubbard band are well separated for $U\gg W$. One might expect that the optical absorption for large interactions, $U \gg W$, and high temperatures, $k_{\rm B}T\gg J={\cal O}(W^2/U)$, shows the signature of a broad band-to-band transition for $U-W\leq \omega\leq U+W$ (units $\hbar\equiv 1$), similar to the Peierls insulator (Gebhard {\em et al.} I 1996). Such considerations seemed to be supported by analytical (Lyo {\em et al.} 1977); (Lyo 1978); (Galinar 1979) and numerical calculations (Campbell, Gammel, and Loh 1989). Below we will calculate~$\sigma(\omega>0)$ in the limit $U\gg W$, and show that the linear absorption is actually dominated by a singular contribution at~$\omega=U$ because the upper and lower Hubbard band are in fact {\em parallel\/} bands. The situation changes for $k_{\rm B}T\ll J$ which we will consider in (Gebhard, Bott, Scheidler, Thomas, and Koch III 1996). \subsection{Harris-Lange model} In the following we will address the limit $W/U\to 0$ where matters considerably simplify since the charge and spin degrees of freedom completely decouple (Ogata and Shiba 1990); (Parola and Sorella 1990). For example, for less than half-filling, $N \leq L$, and $U=\infty$ the eigenenergies become those of a Fermi gas of $N_h=L-N$ holes with dispersion~$\epsilon(k)$, and each energy level is $2^N$-fold degenerate in the thermodynamical limit (Beni, Pincus, and Holstein 1973); (Klein 1973); (Ogata {\em et al.} 1990); (Parola {\em et al.} 1990). To facilitate the discussion of the strong coupling limit we map the Hubbard model onto a problem for which the number of double occupancies is {\em conserved}. For a large on-site Coulomb repulsion $W/U \to 0$ it is natural to start with a spectral decomposition of operators into those which solely act in the upper or lower Hubbard band, and to perturbatively eliminate those parts in $\hat{H}_{\rm Hubbard}$ which couple the two bands. For the Hubbard model this has first been achieved by Harris and Lange (Harris and Lange 1967); (van Dongen 1994), and the resulting effective Hamiltonian to lowest order in~$W/U$ will thus be called the ``Harris-Lange'' model. It offers several advantages, both for analytical and numerical calculations. To carry out the spectral decomposition we start from the case $t=0$. The Fermi annihilation operator can be split into a part which destroys an electron on a single occupied site and does not change the energy of the state, and another part which destroys an electron on a double occupied site and thus decreases the energy by $U$, \begin{equation} \hat{c}_{l,\sigma} = \hat{n}_{l,-\sigma} \hat{c}_{l,\sigma} + (1 - \hat{n}_{l,-\sigma}) \hat{c}_{l,\sigma} \quad . \end{equation} The corresponding creation operator can be treated accordingly. If we now turn on the hopping of electrons ($t\neq 0$) we may split the kinetic energy operator into \begin{mathletters} \label{HLM} \begin{eqnarray} \hat{T} &=& \hat{T}_{\rm LHB} + \hat{T}_{\rm UHB} + \hat{T}^+ + \hat{T}^- \\[6pt] \hat{T}_{\rm LHB} &=& (-t) \sum_{l,\sigma} \left(1-\hat{n}_{l,-\sigma}\right) \left( \hat{c}_{l,\sigma}^+ \hat{c}_{l+1,\sigma}^{\phantom{+}} + \hat{c}_{l+1,\sigma}^+ \hat{c}_{l,\sigma}^{\phantom{+}} \right) \left( 1-\hat{n}_{l+1,-\sigma}\right) \\[6pt] \hat{T}_{\rm UHB} &=& (-t) \sum_{l,\sigma} \hat{n}_{l,-\sigma} \left( \hat{c}_{l,\sigma}^+ \hat{c}_{l+1,\sigma}^{\phantom{+}} + \hat{c}_{l+1,\sigma}^+ \hat{c}_{l,\sigma}^{\phantom{+}} \right) \hat{n}_{l+1,-\sigma}\\[6pt] \hat{T}^+ &=& (-t) \sum_{l,\sigma} \left[ \hat{n}_{l,-\sigma} \hat{c}_{l,\sigma}^+ \hat{c}_{l+1,\sigma}^{\phantom{+}} \left( 1-\hat{n}_{l+1,-\sigma} \right) + \hat{n}_{l+1,-\sigma} \hat{c}_{l+1,\sigma}^+ \hat{c}_{l,\sigma}^{\phantom{+}} \left( 1- \hat{n}_{l,-\sigma}\right) \right] \\[6pt] \hat{T}^- &=& \left(\hat{T}^+\right)^+ \; . \end{eqnarray} \end{mathletters}% The operator~$\hat{T}_{\rm LHB}$ for the lower Hubbard band describes the hopping of holes while doubly occupied sites can move in the upper Hubbard band via $\hat{T}_{\rm UHB}$. Their number is conserved by both hopping processes. These two bands will constitute the basis for our approach. The operator~$\hat{T}^+$ ($\hat{T}^-$) increases (decreases) the number of double occupancies by one. Similar to the Foldy-Wouthuysen transformation for the Dirac equation (Bjorken and Drell 1964) we apply a canonical transformation that eliminates the operators~$\hat{T}^{\pm}$ to a given order in~$t/U$, \begin{equation} \hat{c}_{l,\sigma} = e^{i\hat{S}(\bar{c})} \bar{c}_{l,\sigma} e^{-i\hat{S}(\bar{c})} \end{equation} with $\left(\hat{S}(\bar{c})\right)^+=\hat{S}(\bar{c})$. As shown by Harris and Lange (Harris {\em et al.} 1967); (van Dongen 1994) the operator to lowest order in~$t/U$ reads \begin{equation} \hat{S}(\bar{c}) = \frac{it}{U} \left( \hat{T}^{\bar{c},+} -\hat{T}^{\bar{c},-} \right) \end{equation} which can easily be verified since $\left[\hat{D},\hat{T}^{\pm}\right]_{-}= \pm \hat{T}^{\pm}$. The transformed Hamilton operator in the new Fermions becomes the Harris-Lange model \begin{equation} \hat{H}_{\rm HL}^{\bar{c}} = \hat{T}_{\rm LHB}^{\bar{c}} + \hat{T}_{\rm UHB}^{\bar{c}} + U \hat{D}^{\bar{c}} \quad , \end{equation} if we neglect all correction terms to order~$t/U$ and higher. The energies obtained from the Harris-Lange model thus agree with those of the Hubbard model to order~$t (t/U)^{-1}$ and~$t (t/U)^0$. For all other physical operators which do not contain a factor of~$U/t$ we may replace \begin{equation} \hat{c}_{l,\sigma} = \bar{c}_{l,\sigma} \quad . \end{equation} because the error is only of order $\left(t/U\right)$. In the following we will thus make no distinction between the operators $\hat{c}_{l,\sigma}$ and $\bar{c}_{l,\sigma}$ to lowest order in~$t/U$. The Hamiltonian has the following symmetry. The particle-hole transformation \begin{mathletters} \label{parthole} \begin{eqnarray} {\cal T}_{\rm ph} \hat{c}_{l,\sigma}^+ {\cal T}_{\rm ph}^{-1}& = & i \lambda_{\sigma} e^{i\pi l} \hat{c}_{l,-\sigma}^{\phantom{+}} \\[6pt] {\cal T}_{\rm ph} \hat{c}_{k,\sigma}^+ {\cal T}_{\rm ph}^{-1}& = & i\lambda_{\sigma} \hat{c}_{\pi/a-k,-\sigma}^{\phantom{+}} \label{watchout} \end{eqnarray} with $\lambda_{\uparrow}=-\lambda_{\downarrow}=1$ is generated with the help of \begin{eqnarray} {\cal T}_{\rm ph} &=& e^{i\pi/2 (\hat{C}^+ +\hat{C}^-)}= \prod_l \left[ 1-(\hat{D}_l+\hat{H}_l) + i(\hat{C}_l^+ +\hat{C}_l^-)\right] \\[6pt] \hat{C}^+&=&\left(\hat{C}^-\right)^+=\sum_l \hat{C}_l^+= \sum_l (-1)^l \hat{c}_{l,\uparrow}^+ \hat{c}_{l,\downarrow}^+ \quad \quad;\quad \hat{H}_l=(1-\hat{n}_{l,\uparrow})(1-\hat{n}_{l,\downarrow}) \; . \label{defofthecs} \end{eqnarray} \end{mathletters}% The additional phase factors $i\lambda_\sigma$ are irrelevant global phases, and can be ignored since there is always an equal number of Fermion creation and annihilation operators of each spin species. The operators for the motion of holes and double occupancies are mapped into each other, \begin{equation} \hat{T}_{\rm UHB} \mapsto \hat{T}_{\rm LHB} \qquad \hat{T}_{\rm LHB} \mapsto \hat{T}_{\rm UHB} \; . \end{equation} Furthermore, $\left[ \hat{T}_{\rm UHB}+\hat{T}_{\rm LHB}, \hat{C}^{\pm}\right]_-=0$. This symmetry allows for an exact solution of the model since there is essentially no difference in the motion of double occupancies in the upper Hubbard band and holes in the lower Hubbard band. The discussion above is readily generalized to the case of dimerization in the Harris-Lange model. The model Hamiltonian reads \begin{equation} \hat{H}_{\rm HL}^{\rm dim} = \hat{T}_{\rm LHB}(\delta) + \hat{T}_{\rm UHB}(\delta) + U\hat{D} \end{equation} in an obvious generalization of the kinetic operators for the upper and lower Hubbard bands. \subsection{Optical absorption and optical conductivity} The dielectric function~$\widetilde{\epsilon}(\omega)$ and the coefficient for the linear optical absorption~$\widetilde{\alpha}(\omega)$ are given by (Haug and Koch 1990) \begin{mathletters} \begin{eqnarray} \widetilde{\epsilon}(\omega) &=& 1 +\frac{4\pi i \sigma(\omega)}{\omega} \label{epssigma}\\[6pt] \widetilde{\alpha}(\omega) &=& \frac{4\pi {\rm Re}\{\sigma(\omega)\}}{n_b c} \end{eqnarray} \end{mathletters}% where ${\rm Re}\{\ldots\}$ denotes the real part and $n_b$ is the background refractive index. It is supposed to be frequency independent near a resonance. Hence, the real part of the optical conductivity directly gives the absorption spectrum of the system. The standard result (Maldague 1977); (Mahan 1990) for the real part of the optical conductivity in terms of the current-current correlation function~$\chi(\omega)$ is \begin{eqnarray} {\rm Re}\{ \sigma(\omega) \} &=&\frac{{\rm Im}\{\chi(\omega)\}}{\omega} \\[6pt] \chi(\omega) & =& \frac{{\cal N}_{\perp}}{La} i \int_0^{\infty} dt e^{i\omega t} \langle \left[\hat{\jmath}(t),\hat{\jmath}\right]_- \rangle \end{eqnarray} where~${\cal N}_{\perp}$ is the number of chains per unit area perpendicular to the chain direction. The current-current correlation function can be spectrally decomposed in terms of exact eigenstates of the system as \begin{equation} \chi(\omega) = \frac{{\cal N}_{\perp}}{La} \sum_n |\langle 0 | \hat{\jmath} | n\rangle|^2 \left[ \frac{1}{\omega +(E_n-E_0) +i\gamma} - \frac{1}{\omega -(E_n-E_0) +i\gamma} \right] \; . \label{decomp} \end{equation} Here, $|0\rangle$ is the exact ground state (energy $E_0$), $|n\rangle$ are exact excited states (energy $E_n$), and $\left|\langle 0 | \hat{\jmath} | n\rangle\right|^2$ are the oscillator strengths for optical transitions between them. Although $\gamma =0^+$ is infinitesimal we may introduce $\gamma>0$ as a phenomenological broadening of the resonances at $\omega = \pm(E_n-E_0)$. The spectral decomposition of the real part of the optical conductivity reads \begin{equation} {\rm Re}\{ \sigma(\omega) \} = \frac{{\cal N}_{\perp} \pi}{La \omega} \sum_n\left| \langle 0 | \hat{\jmath} | n\rangle\right|^2 \left[ \delta\left( \omega -(E_n-E_0)\right) -\delta\left( \omega +(E_n-E_0)\right) \right] \label{speccomp} \end{equation} which is positive for all~$\omega$. In the following we will always plot the dimensionless reduced optical conductivity \begin{equation} \sigma_{\rm red}(\omega >0) = \frac{\omega {\rm Re}\{\sigma(\omega>0)\} }% {{\cal N}_{\perp}a e^2 W } \; . \label{sigmared} \end{equation} Furthermore we replace the energy conservation~$\delta(x)$ by the smeared function \begin{equation} \widetilde{\delta}(x) = \frac{\gamma}{\pi(x^2+\gamma^2)} \end{equation} to include effects of phonons, and experimental resolution. For all figures we graphically checked that the sum rules of appendix~\ref{appsumrule} were fulfilled. \subsection{Current operator} As derived in (Gebhard {\em et al.} I 1996) the current operator is given by \begin{equation} \hat{\jmath} = -e \sum_{l,\sigma} ita \left(1+ (-1)^l\delta\right)\left(1+ (-1)^l\eta\right) \left( \hat{c}_{l+1,\sigma}^+ \hat{c}_{l,\sigma}^{\phantom{+}} - \hat{c}_{l,\sigma}^+ \hat{c}_{l+1,\sigma}^{\phantom{+}} \right) \label{currenta} \end{equation} where $\eta=-|R_{l+1}-R_l-a|/a<0$ is the relative change of lattice distances due to the Peierls distortion. Note that $\delta$ and $\eta$ always have opposite sign. The current operator can be split into two parts, $\hat{\jmath}= \hat{\jmath}_{\rm intra}^{\rm H} +\hat{\jmath}_{\rm inter}^{\rm H}$, where $\hat{\jmath}_{\rm intra}^{\rm H}$ moves electrons between neighboring sites without changing the number of double occupancies or holes. This (Hubbard-)intraband current does not change the number of double occupancies. Hence it can be ignored for the optical absorption in the Harris-Lange model at half-filling. The current operator between the two Hubbard bands $\hat{\jmath}_{\rm inter}^{\rm H}$ can be written as \begin{mathletters} \begin{eqnarray} \hat{\jmath}_{\rm inter}^{\rm H} &=& \hat{\jmath}_{{\rm inter},+}^{\rm H} + \hat{\jmath}_{{\rm inter},-}^{\rm H} \\[6pt] \hat{\jmath}_{{\rm inter},+}^{\rm H} &=& -(itea) \sum_{l,\sigma}\left(1+(-1)^l \delta\right) \left(1+(-1)^l \eta\right) \nonumber \\[3pt] && \phantom{-(itea)\sum} \left[ \hat{n}_{l+1,-\sigma} \hat{c}_{l+1,\sigma}^+ \hat{c}_{l,\sigma}^{\phantom{+}} \left( 1- \hat{n}_{l,-\sigma} \right) - \hat{n}_{l,-\sigma} \hat{c}_{l,\sigma}^+ \hat{c}_{l+1,\sigma}^{\phantom{+}} \left( 1- \hat{n}_{l+1,-\sigma} \right) \right] \\[6pt] \hat{\jmath}_{{\rm inter},-}^{\rm H} &=& -(itea) \sum_{l,\sigma}\left(1+(-1)^l \delta\right) \left(1+(-1)^l \eta\right) \nonumber \\[3pt] && \phantom{-(itea)\sum} \left[ \left(1- \hat{n}_{l+1,-\sigma}\right) \hat{c}_{l+1,\sigma}^+ \hat{c}_{l,\sigma}^{\phantom{+}} \hat{n}_{l,-\sigma} - \left(1 -\hat{n}_{l,-\sigma} \right) \hat{c}_{l,\sigma}^+ \hat{c}_{l+1,\sigma}^{\phantom{+}} \hat{n}_{l+1,-\sigma} \right] \end{eqnarray} \end{mathletters}% where~$\hat{\jmath}_{\rm inter,\pm}^{\rm H}$ create and destroy a neighboring pair of double occupancy and hole, respectively. Next we study the action of $\hat{\jmath}_{{\rm inter},+}^{\rm H}$ on a pair of neighboring spins in a state~$|\Psi\rangle$ in position space. It is a sequence of singly occupied sites ($\sigma$), holes ($\circ$), and double occupancies ($\bullet$) from site $1$ to~$L$, e.g., \begin{equation} |\Psi\rangle = |\uparrow_1, \bullet_2, \circ_3,\circ_4,\downarrow_5, \ldots\uparrow_{L-3},\bullet_{L-2},\downarrow_{L-1},\bullet_{L}\rangle \quad . \label{defofstate} \end{equation} We introduce the notations \begin{mathletters} \begin{eqnarray} |\ldots,(\uparrow_l,\downarrow_{l+1}\pm \downarrow_{l},\uparrow_{l+1}), \ldots\rangle &=& |\ldots, \uparrow_l, \downarrow_{l+1},\ldots \rangle \pm |\ldots,\downarrow_{l},\uparrow_{l+1},\ldots\rangle\\[3pt] |{\rm S}_{l,l+1}=1,S^z_{l,l+1}=1\rangle&=& |\ldots,\uparrow_l,\uparrow_{l+1}, \ldots\rangle\\[3pt] |{\rm S}_{l,l+1}=1,S^z_{l,l+1} =0\rangle &=& |\ldots,(\uparrow_l,\downarrow_{l+1} +\downarrow_{l},\uparrow_{l+1}),\ldots\rangle\\[3pt] |{\rm S}_{l,l+1}=1,S^z_{l,l+1}=-1\rangle &=& | \ldots,\downarrow_l,\downarrow_{l+1}, \ldots\rangle\\[3pt] |{\rm S}_{l,l+1}=0,S^z_{l,l+1} =0\rangle &=& |\ldots,(\uparrow_l,\downarrow_{l+1} -\downarrow_{l},\uparrow_{l+1}),\ldots\rangle \end{eqnarray} as the local spin triplet and spin singlet states. Furthermore, \begin{equation} |C_{l,l+1}=1,C^z_{l,l+1}=0\rangle = |\ldots,(\bullet_{l},\circ_{l+1} - \circ_l,\bullet_{l+1}),\ldots\rangle \end{equation} \end{mathletters}% denotes the local charge triplet state since $\hat{C}^+|C_{l,l+1}=1,C^z_{l,l+1}=0\rangle \neq 0$. With these definitions one finds \begin{mathletters} \begin{eqnarray} \hat{\jmath}_{\rm inter,+}^{\rm H} |{\rm S}_{l,l+1}=1\rangle &=&0 \\[6pt] \hat{\jmath}_{\rm inter,+}^{\rm H} |{\rm S}_{l,l+1}=0\rangle &=& -itea (1+(-1)^l\delta)(1+(-1)^l\eta)(-2) |C_{l,l+1}=1,C^z_{l,l+1}=0\rangle \; . \end{eqnarray} \end{mathletters}% It is thus seen that $\hat{\jmath}_{\rm inter}^{\rm H}$ preserves the spin of a neighboring pair such that $\Delta S=\Delta S^z=0$ is the selection rule for the spin sector. The selection rule for the charge sector is found as $\Delta C=1$, $\Delta C^z=0$. Note that the current operator does {\em not\/} commute with~$\hat{C}^{\rm \pm}$ as defined in eq.~(\ref{defofthecs}). Finally, the current operator is invariant against translations by one unit cell and thus preserves the total momentum modulo a reciprocal lattice vector ($Q=2\pi/a$ for~$\delta=0$, $Q=\pi/a$ for~$\delta\neq 0$). However, the current operator can create or destroy a charge excitation with momentum~$q$, and create or destroy a spin excitation with momentum~$-q$. Although there is charge-spin separation in the Hubbard model for strong coupling the current operator mixes both degrees of freedom. This renders the calculation of the optical absorption of the Hubbard model a very difficult problem even in the limit of strong correlations. \section{Exact solution of the Harris-Lange model} \label{exactsolution} \subsection{Translational invariant case} The Harris-Lange model can exactly be solved by an explicit construction of all eigenstates. This has recently been shown by (de Boer, Korepin, and Schadschneider 1995) and (Schadschneider 1995) for periodic, and by (Aligia and Arrachea 1994) for open boundary conditions. Since optical excitations conserve total momentum we work with periodic boundary conditions where the total momentum is a good quantum number. The number~$N_S$ of sites with spin (singly occupied sites), and the number~$N_C=N_d+N_h$ of sites with charge (double occupancies and holes) are separately conserved in the Harris-Lange model. We have $N_C+N_S=L$ lattice sites, $N=N_S+2N_d$ electrons, and choose~$L$ to be even such that our lattice is bipartite for all~$L$. In the sequence of singly occupied sites, double occupancies and holes of the state~$|\Psi\rangle$ in eq.~(\ref{defofstate}) we may identify subsequences for the spins and the charges only (independent of the position on a special site). Additionally, the indices $l_j$ indicate the actual position of the charges $C_j$. The positions occupied by the spins are then the ones left over by the charges: \begin{mathletters} \begin{equation} |\Psi\rangle= | ({l_1},{l_2}, \ldots{l_{N_{C}-1}},{l_{N_C}}); (C_{1},C_{2},\ldots C_{N_{C}-1},C_{N_C}); (S_{1},S_{2},\ldots S_{N_{S}-1},S_{N_S})\rangle \end{equation} where in our example \begin{equation} (S_{1},S_{2},\ldots S_{N_{S}-1},S_{N_S})= (\uparrow,\downarrow,\ldots\uparrow,\downarrow) \end{equation} is the subsequence for the spins and \begin{equation} (C_{1},C_{2},C_{3},\ldots C_{N_{C}-1},C_{N_C})= (\bullet,\circ,\circ,\ldots\bullet,\bullet) \end{equation} for the charges. The sequence for the positions occupied by charges is \begin{equation} ({l_1},{l_2},{l_3},\ldots {l_{N_C}-1},{l_{N_C}})=(2,3,4,\ldots L-2,L) \qquad . \end{equation} \end{mathletters}% Since there is nearest-neighbor hopping only both the spin and charge subsequences are separately {\em conserved\/} up to cyclic permutations due to the periodic boundary conditions. To include the boundary effect we follow (de Boer {\em et al.} 1995) and (Schadschneider 1995) and introduce the properly symmetrized spin and charge sequences. To this end we define the operator for a cyclic permutation of the spin sequence, \begin{equation} \hat{\cal T}_S (S_1,S_2,\ldots S_{N_S}) = (S_{N_S},S_1,\ldots S_{N_S-1}) \quad , \end{equation} and, equivalently, $\hat{\cal T}_C$ for the charge sequence. Let $K_S$ and $K_C$ be the smallest positive integers such that $\left(\hat{\cal T}_S\right)^{K_S}$ and $\left(\hat{\cal T}_C\right)^{K_C}$ act as identity operators on a given spin and charge sequence, respectively. Then we define the~$K_S K_C$ states {\arraycolsep=0pt\begin{eqnarray} &&\sqrt{K_SK_C} |({l_1}, \ldots {l_{N_C}}); (C_1,\ldots C_{N_C})_{k_C}; (S_1,\ldots S_{N_S})_{k_S}\rangle = \label{realspace}\\[9pt] && e^{\sum_l i\pi l \hat{D}_{l}} \sum_{\nu_S=0}^{K_S-1} \sum_{\nu_C=0}^{K_C-1} e^{ia(\nu_S k_S+\nu_C k_C)} \hat{\cal T}_S^{\nu_S} \hat{\cal T}_C^{\nu_C} |({l_1}, \ldots {l_{N_C}}); (C_1,\ldots C_{N_C});(S_1,\ldots S_{N_S})\rangle \nonumber \end{eqnarray}}% with the momentum shifts $k_S=2\pi m_S/(K_S a)$, $m_S=0,1,\ldots (K_S-1)$, $k_C=2\pi m_C/(K_C a)$, $m_C=0,1,\ldots (K_C-1)$. An extra phase factor~$(-1)^{l_j}$ for each double occupancy at site~$l_j$ has been included here through the operator $\exp(\sum_l i\pi l \hat{D}_{l})$. This allows to make direct contact to the model considered by (de Boer {\em et al.} 1995); (Schadschneider 1995), and (Aligia {\em et al.} 1994). There the hopping amplitudes for lower and upper Hubbard band have opposite signs. We transform the states of eq.~(\ref{realspace}) into momentum space with respect to the charge coordinates. The exact eigenstates can now be classified according to their spin and charge sequence, their momentum shifts $k_C$, $k_S$, and~$N_C$ momenta from the set of~$k_{j}=2\pi m_j/(L a)$, $m_j=-(L/2),\ldots (L/2)-1$. The normalized eigenstates read {\arraycolsep=0pt\begin{eqnarray} &&L^{N_C/2}|k_1,\ldots k_{N_C}; (C_1,\ldots C_{N_C})_{k_C}; (S_1,\ldots S_{N_S})_{k_S}\rangle = \label{eigenstates} \\[9pt] &&\sum_{l_1< \ldots <l_{N_C}} \!\! \sum_{{\cal P}}(-1)^{\cal P} \exp \! \left(ia\sum_{j=1}^{N_C} l_{{\cal P}(j)}\left(k_j+\Phi_{CS}\right) \! \right) \! |({l_1}, \ldots {l_{N_C}}); (C_1,\ldots C_{N_C})_{k_C}; (S_1,\ldots S_{N_S})_{k_S}\rangle \nonumber \end{eqnarray}}% where the permutations ${\cal P}$ generate a simple Slater determinant for the momenta and the positions of the $N_C$ charges, and $\Phi_{CS}=(k_C-k_S)/L$ is an additional momentum shift which vanishes in the thermodynamical limit. It is straightforward (de Boer {\em et al.} 1995); (Schadschneider 1995) but lengthy to explicitly show that the states in eq.~(\ref{eigenstates}) have energy and momentum \begin{mathletters} \begin{eqnarray} E&=&\sum_{j=1}^{N_C} \epsilon(k_j+\Phi_{CS}) +U N_d \label{eigenenHL}\\[3pt] P&=& k_S +(\pi/a) (N_d-1)+ \sum_{j=1}^{N_C} (k_j+\Phi_{CS}) + (N {\rm \ mod\ } 2) \pi/a \qquad {\rm mod\ }2\pi/a \label{eigenmomHL} \end{eqnarray} \end{mathletters}% with $\epsilon(k)$ given by equation~(\ref{gl3a}). The essential arguments are repeated in the appendices~\ref{appmomentum} and~\ref{appenerg}. The above set of eigenstates is complete. After summing over the subspaces with different~$K_S$, $K_C$ the number of states which become degenerate in energy in the thermodynamical limit is given by $2^{N_S} N_C!/(N_d!N_h!)$. The number of possible choices for the momenta is~$L!/(N_C!(L-N_C)!)$ since the momenta are those of a gas of spinless Fermions. Altogether one finds for an even number of electrons~$N$ \begin{equation} \sum_{N_d=0}^{N/2} {N_C \choose N_d} 2^{N-2N_d} {L \choose {L-N+2N_d}} = {2L \choose N} \end{equation} where eq.~(10.33.5) of (Hansen 1975) and eq.~(22.2.3) of (Abramovitz and Stegun 1970) have been used. This exhausts the Hilbert space for fixed number of electrons~$N$. \subsection{Dimerized Harris-Lange model} The Harris-Lange model can also be solved in the presence of a finite lattice distortion as long as there is hopping between nearest neighbors only. For the $U=\infty$ Hubbard model at less than half filling this has already been realized some time ago (Bernasconi, Rice, Schneider, and Str\"{a}\ss ler 1975). The exact eigenenergies are those of spinless Fermions on a dimerized chain. For the Harris-Lange model with~$N_C$ charge excitations we choose momenta~$|k_j|\leq \pi/(2a)$ of the reduced Brillouin zone, and one of the~$2^{N_C}$ sequences $(\tau_1,\ldots \tau_{N_C})$ with~$\tau_j=\pm 1$. Let us introduce the operator~$\hat{\Pi}_r$ by {\arraycolsep=0pt\begin{eqnarray} \hat{\Pi}_r |k_1,\ldots k_r,\ldots k_{N_C}; && (C_1,\ldots C_{N_C})_{k_C}; (S_1,\ldots S_{N_S})_{k_S}\rangle \nonumber \\[6pt] && = |k_1,\ldots k_r+\pi/a,\ldots k_{N_C}; (C_1,\ldots C_{N_C})_{k_C}; (S_1,\ldots S_{N_S})_{k_S}\rangle \end{eqnarray}}% and the four functions~$\xi_1^r(1)=\beta_{k_r}$, $\xi_2^r(1)=i\alpha_{k_r}$, $\xi_1^r(-1)=\alpha_{k_r}$, and $\xi_2^r(-1)=-i\beta_{k_r}$, compare eq.~(\ref{alphabeta}). The eigenstates for fixed number of charge excitations~$N_C$ can then be written as \begin{eqnarray} |k_1,\tau_1;\ldots &&k_{N_C},\tau_{N_C}; (C_1,\ldots C_{N_C})_{k_C}; (S_1,\ldots S_{N_S})_{k_S}\rangle = \nonumber \\[6pt] &&\left[ \prod_{r=1}^{N_C} \left( \xi_1^r(\tau_r) + \xi_2^r(\tau_r)\hat{\Pi}_r\right) \right] |k_1,\ldots k_{N_C}; (C_1,\ldots C_{N_C})_{k_C}; (S_1,\ldots S_{N_S})_{k_S}\rangle \; . \end{eqnarray} This state corresponds to $(N_C+\sum_r\tau_r)$ ($(N_C-\sum_r\tau_r)/2)$ spinless Fermions in the upper (lower) Peierls subband. The corresponding energies and momenta of these states are obviously given by \begin{mathletters} \begin{eqnarray} E&=&\sum_{j=1}^{N_C} E(k_j+\Phi_{CS})\tau_j +U N_d\\[3pt] P&=& k_S + \sum_{j=1}^{N_C} (k_j+\Phi_{CS}) \quad {\rm mod\ }\pi/a \end{eqnarray} \end{mathletters}% where~$E(k)$ has been given in eq.~(\ref{peierlsen}). \subsection{Band picture interpretation of the spectrum} \label{bandpictureinterpretation} \subsubsection{Translational invariant case} The exact solution for the Harris-Lange model can be interpreted in terms of upper and lower Hubbard bands. To simplify the discussion we will ignore the momentum shift~$\Phi_{CS}$ in this subsection. For $U>W$ the ground state of the half-filled band~$N=L$ has energy zero and is $2^L$-fold degenerate, and we may choose the fully polarized ferromagnetic state as our reference state, $|{\rm FM}\rangle=|\uparrow,\ldots \uparrow\rangle$. Note that this state has momentum~$\pi/a$ on a chain with an even number of sites~$L$, see eq.~(\ref{eigenmomHL}). We may now add an electron. We obtain all exact eigenstates for $N=L+1$ electrons, $N_d=1$, and all spins up as \begin{mathletters} \begin{equation} |k;(\bullet)_{k_C=0},(\uparrow,\ldots \uparrow)_{k_S=0}\rangle = \hat{c}_{k,\downarrow}^+ |{\rm FM}\rangle \label{plusone} \end{equation} with momentum~$P=k+\pi/a$ and energy~$E=\epsilon(k)+U$. The state in eq.~(\ref{plusone}) is interpreted as a {\em particle\/} at momentum~$k$ in the upper Hubbard band which has the dispersion relation~$\epsilon(k)+U$. The momentum~$\pi/a$ is attributed to the ferromagnetic reference state. We may equally well take out an electron from the fully polarized state. We obtain all exact eigenstates for $N=L-1$ electrons, $N_h=1$, and all spins up as \begin{equation} |k;(\circ)_{k_C=0},(\uparrow,\ldots \uparrow)_{k_S=0}\rangle = - \hat{c}_{\pi/a-k,\uparrow} |{\rm FM}\rangle \label{minusone} \end{equation} \end{mathletters}% with momentum~$P=-(\pi/a-k)+\pi/a$ and energy~$E=\epsilon(k)$. Note that the states of eq.~(\ref{minusone}) and those of eq.~(\ref{plusone}) can be generated from each other by the particle-hole transformation of eq.~(\ref{parthole}). Their momenta differ by~$\pi/a$ since their respective numbers of double occupancies differ by one. Since the ground state corresponds to a completely filled lower Hubbard band we {\em interpret\/} the state in eq.~(\ref{minusone}) as a hole in the lower Hubbard band at~$k_h=\pi/a-k$, and the momentum~$\pi/a$ is again attributed to the ferromagnetic reference state. The lower Hubbard band must have the dispersion relation~$\epsilon(k)$ for {\em particles\/} because a {\em hole\/} at $k_h=\pi/a-k$ has momentum~$P=-(\pi/a-k)$ and energy~$E=-\epsilon(k_h)= -\epsilon(-k+\pi/a)=\epsilon(k)$. The band structure for the Harris-Lange model is depicted in figure~\ref{HarrisLangedis}. It displays the parallel upper and lower Hubbard bands with band width~$W$ separated by a distance~$U$. It is amusing that the celebrated Hubbard-I approximation (Hubbard 1963); (Mazumdar and Soos 1981) also gives parallel bands. Those bands, however, carry a spin index while charge-spin separation is most essential in one dimension. Furthermore, the width of those bands is only {\em half\/} of the exact band width~$W$. Our band structure picture has to be used carefully if there are more than one double occupancy or hole. Figure~\ref{HarrisLangedis} suggests that there are~$L$ states available {\em both\/} in the upper {\em and\/} in the lower Hubbard band, altogether~$2L$ independent states. However, this cannot be the case because for~$N_d=N_h=L/2$ we would have~${L \choose L/2}{L \choose L/2}$ states in the band picture while the Hilbert space actually has only the dimension~${L \choose L/2}$. The exact solution shows how an appropriate exclusion principle between particles in the upper Hubbard band and holes in the lower Hubbard band can be formulated. For fixed spin background and fixed $k_1$, $k_2$ there are four exact eigenstates with two charges at~$k_1\neq k_2$. They all have the kinetic energy~$T=\epsilon(k_1)+\epsilon(k_2)$. They correspond to four different charge excitations in the band picture: (i)~two particles at momenta~$k_1$, $k_2$ in the upper Hubbard band, (ii)~two holes at momenta~$\pi/a-k_1$, $\pi/a-k_2$ in the lower Hubbard band, (iii)~a particle at momentum~$k_1$ in the upper Hubbard band and a hole at momentum~$\pi/a-k_2$ in the lower Hubbard band, (iv)~a particle at momentum~$k_2$ in the upper Hubbard band and a hole at momentum~$\pi/a-k_1$ in the lower Hubbard band. The condition~$k_1\neq k_2$ is naturally fulfilled in cases~(i) and~(ii), if we assign a fermionic character to our particles in the upper and holes in the lower Hubbard band, respectively. In case~(iii), however, we have to explicitly {\em demand\/} that the momentum at which we create the hole, $k_h=\pi/a-k_2$, fulfills~$k_1\neq k_2$, i.~e., $k_h\neq \pi/a-k_1$. This is the same condition which results from case~(iv). We thus see that a particle in the upper Hubbard band at momentum~$k$ actually blocks the creation of a hole in the lower Hubbard band at momentum~$\pi/a-k$ (this is probably the simplest example of a ``statistical interaction'', see (Haldane 1991)). With this additional rule the counting of states in the band picture is correct, and the band picture interpretation gives indeed the {\em exact\/} results for the Harris-Lange model. The effective Hamiltonian for fermionic particles in the upper ($\hat{u}_k$) and lower ($\hat{l}_k$) Hubbard band thus reads \begin{mathletters} \label{effHL} \begin{eqnarray} \hat{H}_{\rm HL}^{\rm band} &=& \hat{P}_{ul} \sum_{|k|\leq \pi/a}\left[ (U+\epsilon(k)) \hat{n}_{k}^{u} + \epsilon(k) \hat{n}_{k}^{l} \right] \hat{P}_{ul} \\[6pt] \hat{P}_{ul} &=& \prod_{|k|\leq \pi/a} \left[ 1 -\left(1- \hat{n}_{\pi/a-k}^{l}\right) \hat{n}_{k}^{u} \right] \end{eqnarray} \end{mathletters}% with $\hat{n}_{k}^{u}=\hat{u}_{k}^+ \hat{u}_{k}^{\phantom{+}}$, $\hat{n}_{k}^{l}=\hat{l}_{k}^+ \hat{l}_{k}^{\phantom{+}}$. The projection operators guarantee that there is no hole in the lower Hubbard band at momentum~$\pi/a-k$, if there is already a particle at momentum~$k$ in the upper Hubbard band. For half-filling at zero temperature the lower Hubbard band is completely filled. \subsubsection{Dimerized Harris-Lange model} The case of the dimerized Hubbard model can be treated accordingly. The upper and lower Hubbard band now split into two Peierls subbands with dispersion relations~$\pm E(k)$. Formally the band structure Hamiltonian for the lower band becomes (compare eq.~(\ref{Tink})) \begin{equation} \hat{T}_{\rm LHB}^{\rm band}(\delta)= \sum_{|k|\leq \pi/(2a)} \epsilon(k)\left( \hat{l}_{k}^+\hat{l}_{k}^{\phantom{+}} - \hat{l}_{k+\pi/a}^+\hat{l}_{k+\pi/a}^{\phantom{+}} \right) -i \Delta(k) \left( \hat{l}_{k+\pi/a}^+\hat{l}_{k}^{\phantom{+}} - \hat{l}_{k}^+\hat{l}_{k+\pi/a}^{\phantom{+}} \right) \; , \label{Tdimnotdiag} \end{equation} and a similar expression holds for the upper Hubbard band. The band picture Hamiltonian can easily be brought into diagonal form as in the Peierls case. The quasi-particles in the four subbands are finally described by \begin{eqnarray} \hat{H}_{\rm HL}^{\rm dim,\, band} &=& \hat{P}_{u^+l^+}\hat{P}_{u^-l^-} \sum_{|k|\leq \pi/(2a)}\biggl[ (U+E(k)) \hat{n}_{k,+}^{u} + (U-E(k)) \hat{n}_{k,-}^{u} \nonumber \\[6pt] && \phantom{\hat{P}_{u^+l^+}\hat{P}_{u^-l^-} \sum_{|k|\leq\pi/(2a)}\biggl[} + E(k) \hat{n}_{k,+}^{l} -E(k) \hat{n}_{k,-}^{l} \biggr] \hat{P}_{u^+l^+}\hat{P}_{u^-l^-} \label{bandhldim} \\[6pt] \hat{P}_{u^{\pm}l^{\pm}} &=& \prod_{|k|\leq\pi/(2a)} \left[ 1- \left(1- \hat{n}_{-k,\pm}^{l}\right)\hat{n}_{k,\pm}^{u}\right] \nonumber \end{eqnarray} with $\hat{n}_{k,\pm}^{u}=\hat{u}_{k,\pm}^+ \hat{u}_{k,\pm}^{\phantom{+}}$, $\hat{n}_{k,\pm}^{l}=\hat{l}_{k,\pm}^+ \hat{l}_{k,\pm}^{\phantom{+}}$ as the number operators for the quasi-particles for the upper ($\tau=+$) and lower ($\tau=-$) Peierls subband in the upper~($u$) and lower~($l$) Hubbard band. The quasi-particles in each subband obey a fermionic exclusion principle in the same Hubbard band. In addition, a particle at momentum~$k$ in the upper Hubbard band in the upper (lower) Peierls subband blocks the creation of a hole at momentum~$-k$ in the lower Hubbard band in the upper (lower) Peierls subband. There is no hole in the lower Hubbard band at momentum~$-k$ in the upper or lower Peierls subband, if there is already a particle at momentum~$k$ in the upper Hubbard band in the corresponding Peierls subband. Note that the reciprocal lattice vector is now given by~$\pi/a$. Thereby the projection operators guarantee the proper counting of states. The resulting band structure is shown in figure~\ref{HueckelHarrisLangedis}. The upper and lower Hubbard band are both Peierls split and display the Peierls gap~$W\delta$ at the zone boundaries~$\pm \pi/(2a)$. Note that the upper (lower) Peierls subbands are still parallel. \subsection{Band picture interpretation of the current operator} \subsubsection{Translational invariant case} According to the spectral decomposition of the current-current correlation function, eq.~(\ref{decomp}), we need to determine the excitation energy~$E_n-E_0$ of an exact eigenstate~$|n\rangle$ and its oscillator strength~$|\langle 0 | \hat{\jmath}|n\rangle|^2$. The respective total momenta of these states are~$P_0$ and~$P_n$. We are interested in optical excitations from a state with singly occupied sites only. The excited states which can be reached from this state have one pair of hole and double occupancy, i.~e., \begin{mathletters} \begin{eqnarray} |0\rangle &=& |(S_1,\ldots S_L)_{k_S}\rangle\\[3pt] |n\rangle &=& |k_1,k_2; (\bullet \circ)_{k_C=0}; (S_1,\ldots S_{L-2})_{k_S^{\prime}}\rangle \end{eqnarray} \end{mathletters}% where we used the fact that~$\hat{\jmath}$ creates a charge triplet with~$k_C=0$. Note that~$(k_1,k_2)$ is the same state as~$(k_2,k_1)$. We denote~$k_1=k+q/2$, $k_2=\pi/a-k+q/2$ since we will finally represent the state~$|n\rangle$ by a particle in the upper Hubbard band at momentum~$k+q/2$ and a hole in the lower Hubbard band at momentum~$k-q/2$. Since~$\hat{\jmath}$ conserves the total momentum we already know that~$P_0=P_n$ which implies $k_S=q+k_S^{\prime}(L-2)/L$, see eq.~(\ref{eigenmomHL}). Hence, $k_S^{\prime}=L(k_S-q)/(L-2)$ has to hold. Recall that~$k_S^{\prime}$ is quantized in units of~$2\pi/((L-2)a)$. These considerations imply that the charge (spin) sector in~$|n\rangle$ carries momentum~$q$ ($-q$) relative to~$|0\rangle$. In the thermodynamical limit the excitation energy is given by \begin{equation} E(k,q)=U + \epsilon(k+q/2) - \epsilon(k-q/2)= U + 4t\sin(ka)\sin(qa/2)\; . \label{EKQ} \end{equation} Note that the excitation energy does not depend on the spin configuration. For this reason it is possible to find a formally equivalent band picture for the charge sector alone that gives the same optical absorption as the original model. Since~$\hat{\jmath}$ itself carries all the information on the conservation laws (momentum, charge, and spin quantum numbers) we may equally well work with the (normalized) states \begin{eqnarray} |k+\frac{q}{2};\frac{\pi}{a}-k+\frac{q}{2}\rangle &=& \frac{1}{L} \sum_{l_1<l_2} \left( e^{i(k+q/2)l_1a}e^{i(\pi/a-k+q/2)l_2a} -e^{i(k+q/2)l_2a}e^{i(\pi/a-k+q/2)l_1a} \right) \nonumber \\ % && \phantom{\frac{1}{L} \sum_{l_1<l_2}} (-1)^{l_1} |S_1^{\prime},\ldots S_{l_1-1}^{\prime},\bullet_{l_1}, S_{l_1}^{\prime},\ldots S_{l_2-2}^{\prime}, \circ_{l_2}, S_{l_2-1}^{\prime},\ldots S_{L-2}^{\prime} \rangle \end{eqnarray} rather than the exact eigenstates of eq.~(\ref{eigenstates}). This will simplify the notation since we do not have to take any summation restrictions into account. For fixed~$(k,q)$ and fixed spin configuration~$(S_1^{\prime},\ldots S_{L-2}^{\prime})$ we calculate \begin{eqnarray} \langle 0 |\hat{\jmath}_{\rm inter,-}^{\rm H} |k+\frac{q}{2};\frac{\pi}{a}-k+\frac{q}{2}\rangle &=& -iea\epsilon(k) e^{iqa/2} \\ && \frac{1}{L} \sum_l e^{iqla} \langle 0 |S_1^{\prime},\ldots S_{l-1}^{\prime}, \left( \uparrow_{l}\downarrow_{l+1}-\downarrow_{l}\uparrow_{l+1}\right), S_{l}^{\prime},\ldots S_{L-2}^{\prime} \rangle \; . \nonumber \end{eqnarray} We define the operators~$\hat{x}_q^+$ and $\hat{x}_q^{\phantom{+}}$ via their product \begin{eqnarray} \hat{x}_q^+ \hat{x}_q^{\phantom{+}} &=& \sum_{S_1^{\prime},\ldots S_{L-2}^{\prime}} \frac{1}{L^2} \sum_{l,r} e^{iq(l-r)a} \langle 0 | S_1^{\prime},\ldots S_{l-1}^{\prime}, \left( \uparrow_{l}\downarrow_{l+1}-\downarrow_{l}\uparrow_{l+1}\right), S_{l}^{\prime},\ldots S_{L-2}^{\prime} \rangle \label{xqxq}\\[6pt] && \phantom{\sum_{S_1^{\prime},\ldots S_{L-2}^{\prime}} \frac{1}{L^2} \sum_{l,r} e^{iq(l-r)} } \langle S_{L-2}^{\prime},\ldots S_{r}^{\prime}, \left( \downarrow_{r+1} \uparrow_{r}-\uparrow_{r+1}\downarrow_{r}\right), S_{r-1}^{\prime},\ldots S_{1}^{\prime} | 0 \rangle \; . \nonumber \end{eqnarray} Summed over all intermediate spin configurations the oscillator strength for fixed $(k,q)$ now becomes \begin{equation} \Bigl|\langle 0 |\hat{\jmath}_{\rm inter,-}^{\rm H} |k+\frac{q}{2};\frac{\pi}{a}-k+\frac{q}{2}\rangle\Bigr|^2 = \Bigl|-iea \epsilon(k)\Bigr|^2 \hat{x}_q^+\hat{x}_q^{\phantom{+}} \; . \end{equation} It is clear that we have hidden a very difficult many-body problem in the operators~$\hat{x}_q$. Nevertheless we are now in the position to identify the interband current operator in the band picture. It is given by \begin{equation} \hat{\jmath}_{\rm inter}^{\rm band}= \sum_{|k|,|q|\leq \pi/a} iea\epsilon(k) \left( \hat{u}_{k+q/2}^+\hat{l}_{k-q/2}^{\phantom{+}} \hat{x}_{q}^{\phantom{+}} - \hat{l}_{k-q/2}^+\hat{u}_{k+q/2}^{\phantom{+}} \hat{x}_{q}^+ \right)\; . \label{jhlbandpicture} \end{equation} This operator acts in the same space as the band Hamiltonian of section~\ref{bandpictureinterpretation}. It is seen that the condition~$k+q/2 \neq \pi/a-(k-q/2)$ is automatically fulfilled since~$\epsilon(\pi/(2a))=0$. Consequently, the projection operators in eq.~(\ref{effHL}) can again be ignored for the case of linear optical absorption. \subsubsection{Dimerized Harris-Lange model} For a distorted lattice the current operator can also modify the momentum of the state by $\pi/a$. Thus we address {\em four\/} possible states for fixed~$(k,q)$ from the reduced Brillouin zone, $|k+q/2;\pi/a-k+q/2\rangle$, $|k+q/2;-k+q/2\rangle$, $|\pi/a+k+q/2;\pi/a-k+q/2\rangle$, and \hbox{$|\pi/a+k+q/2;-k+q/2\rangle$}. The same analysis as in the previous subsection leads us to the definition of the operators~$\hat{x}_q^{+}(\delta,\eta)$, $\hat{x}_q^{\phantom{+}}(\delta,\eta)$ with the property \newpage \typeout{forced newpage} {\arraycolsep=0pt\begin{eqnarray} \hat{x}_q^{+}(\delta,\eta)\hat{x}_{q'}^{\phantom{+}}(\delta,\eta) &=& \sum_{S_1^{\prime},\ldots S_{L-2}^{\prime}} \frac{1}{L^2} \sum_{l,r} e^{i(ql-q'r)a} \bigl(1+\eta\delta +(-1)^l(\delta+\eta)\bigr) \bigl(1+\eta\delta +(-1)^r(\delta+\eta)\bigr) \nonumber \\[6pt] && \phantom{\sum_{S_1^{\prime},\ldots S_{L-2}^{\prime}} \frac{1}{L^2} \sum_{l,r} e^{i(ql-q'r)} } \langle 0 | S_1^{\prime},\ldots S_{l-1}^{\prime}, \left( \uparrow_{l}\downarrow_{l+1}-\downarrow_{l}\uparrow_{l+1}\right), S_{l}^{\prime},\ldots S_{L-2}^{\prime} \label{xqxqprime} \rangle \\[6pt] && \phantom{\sum_{S_1^{\prime},\ldots S_{L-2}^{\prime}} \frac{1}{L^2} \sum_{l,r} e^{i(ql-q'r)} } \langle S_{L-2}^{\prime},\ldots S_{r}^{\prime}, \left( \downarrow_{r+1} \uparrow_{r}-\uparrow_{r+1}\downarrow_{r}\right), S_{r-1}^{\prime},\ldots S_{1}^{\prime} | 0 \rangle \; . \nonumber \end{eqnarray}}% In practice, $q'=q$ or $q'=q+\pi/a$. The interband current operator becomes \begin{eqnarray} \hat{\jmath}_{\rm inter}^{\rm band}&=& \hat{\jmath}_{\rm inter,+}^{\rm band}+ \hat{\jmath}_{\rm inter,-}^{\rm band} \nonumber \\[6pt] \hat{\jmath}_{\rm inter,+}^{\rm band}&=& \left(\hat{\jmath}_{\rm inter,-}^{\rm band}\right)^+ \nonumber \\[6pt] \hat{\jmath}_{\rm inter,-}^{\rm band}&=& \sum_{|q|,|k|\leq \pi/(2a)}\biggl\{ -iea \epsilon(k) \Bigl[ \hat{l}_{k-q/2}^+ \hat{u}_{k+q/2}^{\phantom{+}} - \hat{l}_{k-q/2+\pi/a}^+ \hat{u}_{k+q/2+\pi/a}^{\phantom{+}} \Bigr] \hat{x}_{q}^+ \label{jinterband}\\[9pt] && \phantom{\sum_{|q|,|k|\leq \pi/(2a)}\biggl\{ } +ea \frac{\Delta(k)}{\delta} \Bigl[ \hat{l}_{k-q/2}^+ \hat{u}_{k+q/2+\pi/a}^{\phantom{+}} - \hat{l}_{k-q/2+\pi/a}^+ \hat{u}_{k+q/2}^{\phantom{+}} \Bigr] \hat{x}_{q+\pi/a}^+ \biggr\} \nonumber \; . \end{eqnarray} Again the conditions~$k+q/2 \neq \pi/a-(k-q/2)$ and $k+q/2 \neq -(k-q/2)$ will not be violated since~$\epsilon(\pi/(2a))=0$ and $\Delta(0)=0$, respectively. Consequently, the projection operators in eq.~(\ref{bandhldim}) can be ignored for the case of linear optical absorption. In semiconductor physics one prefers to work with the dipole operator rather than the current operator to set up the perturbation theory in the electrical field (Haug {\em et al.} 1990). The corresponding expressions for the dipole operator for Hubbard interband transitions are derived in appendix~\ref{appdipoleHL}. \section{Optical absorption in the Harris-Lange model} \label{optabsHL} \subsection{Spin average} All states with no double occupancy are possible ground states in the Harris-Lange model at half-filling. Instead of looking at the optical absorption for a specific state~$|0\rangle$ it is more reasonable to calculate the {\em average\/} absorption, i.~e., \begin{equation} \overline{{\rm Im}\{\chi(\omega)\}} = \frac{1}{2^L} \sum_{|0\rangle} {\rm Im}\{\chi_{|0\rangle}(\omega)\} \; . \end{equation} For the Hubbard model this corresponds to temperatures $k_{\rm B}T \gg J={\cal O}(W^2/U)$ (``hot-spin case''). The calculation is performed in appendix~\ref{spiav}. We find the result \begin{mathletters} \begin{eqnarray} \langle \hat{x}_q^+\hat{x}_{q'}\rangle &=& \frac{1}{2L} \biggl\{ \delta_{q,q'} \left[(1+\delta\eta)^2 g(q)+(\delta+\eta)^2 g(q+\frac{\pi}{a})\right] \nonumber \\[3pt] && \phantom{\frac{1}{2L} \biggl\{ } + \delta_{q,q'+\frac{\pi}{a}} (1+\delta\eta)(\delta+\eta) \left[g(q)+ g(q+\frac{\pi}{a})\right] \biggr\}\label{averagexq}\\[3pt] g(q) &=& \frac{3}{5+4\cos(qa)} \; . \end{eqnarray}\end{mathletters}% The spin problem could thus be traced out completely. It is seen that $\hat{x}_q$ keeps its operator character until we have expressed the current operator in terms of the Fermion operators for the four Peierls subbands. \subsection{Optical absorption} \subsubsection{Translational invariant case} The real part of the average optical conductivity becomes \begin{equation} {\rm Re}\{\overline{\sigma(\omega >0)} \} = \frac{\pi {\cal N}_{\perp}}{2 L^2 a \omega} \sum_{|q|,|k|\leq \pi/a} (ea \epsilon(k))^2 g(q) \delta(\omega - E(k,q)) \label{elliptic} \end{equation} with $E(k,q)$ from equation~(\ref{EKQ}). The above expression can be written as \begin{equation} \overline{\sigma_{\rm red}(\omega >0)} = \frac{1}{4\pi}\int_{|u|}^1\frac{dx}{x^2} \frac{\sqrt{x^2-u^2}}{\sqrt{1-x^2}} \frac{3}{9-8x^2} \label{what} \end{equation} for the reduced optical conductivity with $u=|\omega-U|/W\leq 1$. This integral can be expressed as a sum over elliptic integrals but we rather prefer to discuss some special cases. The optical absorption is restricted to $|\omega-U|\leq W$. Near the band edges the absorption increases linearly which can be seen from equation~(\ref{what}) by a transformation~$x\to 1-y/|u|$. The more interesting case is $\omega=U$. Now the integrand displays a $1/x$ singularity for~$|u|\to 0$. The {\em parallel\/} Hubbard bands give rise to a logarithmic divergence, $\sigma(\omega\to U) \sim |\ln(\omega-U)|$, since their large joint density of states for $\omega=U$ survives even in the presence of a spinon bath that provides any momentum to the charge sector. The overall behavior of the optical absorption is shown in figure~\ref{hl00}. The same absorption curve has been obtained earlier in (Lyo {\em et al.} 1977) for their ``random'' spin background. The result for their ``ferromagnetic'' spin background follows when we put~$g(q)\equiv 1$, as expected. We will comment on their N\'{e}el state results in (Gebhard {\em et al.} III 1996). \subsubsection{Dimerized Harris-Lange model} We have to diagonalize the interband current operator of eq.~(\ref{jinterband}) in terms of the Peierls operators for the lower Hubbard band \begin{mathletters} \label{trafoforl} \begin{eqnarray} \hat{l}_{k}&=&\alpha_{k}\hat{l}_{k,-}+\beta_k\hat{l}_{k,+}\\[3pt] \hat{l}_{k+\pi/a}&=&-i\beta_{k}\hat{l}_{k,-}+i\alpha_k\hat{l}_{k,+} \end{eqnarray} \end{mathletters}% for $|k|\leq \pi/(2a)$. The transformation for the upper Hubbard band is analogous. With this definition the Hamiltonian in the band picture interpretation became diagonal, see eq.~(\ref{bandhldim}). The interband current operator becomes \begin{equation} \hat{\jmath}_{\rm inter,-}^{\rm band} =\sum_{\tau,\tau'=\pm 1} \sum_{|k|,|q| \leq \pi/(2a)} \lambda_{\tau,\tau'} (k,q) \hat{l}_{k-q/2,\tau}^+\hat{u}_{k+q/2,\tau'}^{\phantom{+}} \label{jinterdimHL} \end{equation} with \begin{mathletters} \label{thelambdas} \begin{eqnarray} \lambda_{+,+}(k,q) &=& iea \left[ \epsilon(k) (\alpha_{+}^{\phantom{*}}\alpha_{-}^* -\beta_{+}^{\phantom{*}}\beta_{-}^*) \hat{x}_q^+ +\frac{\Delta(k)}{\delta} (\alpha_{+}^{\phantom{*}}\beta_{-}^* +\beta_{+}^{\phantom{*}}\alpha_{-}^*) \hat{x}_{q+\pi/a}^+ \right] \\[6pt] \lambda_{+,-}(k,q) &=& iea \left[ - \epsilon(k) (\alpha_{+}^{\phantom{*}}\beta_{-}^* +\beta_{+}^{\phantom{*}}\alpha_{-}^*) \hat{x}_q^+ +\frac{\Delta(k)}{\delta} (\alpha_{+}^{\phantom{*}}\alpha_{-}^* -\beta_{+}^{\phantom{*}}\beta_{-}^*) \hat{x}_{q+\pi/a}^+ \right] \end{eqnarray} \end{mathletters}% and $\lambda_{-,-}(k,q)= -\lambda_{+,+}(k,q)$, $\lambda_{-,+}(k,q)=\lambda_{+,-}(k,q)$. Here we used the short-hand notation $\alpha_{\pm}=\alpha_{k\pm q/2}$ etc. Note that these quantities can be complex for~$q \neq 0$. The average optical conductivity becomes \begin{mathletters} \label{monstersigma} \begin{equation} {\rm Re}\{\overline{\sigma(\omega >0, \delta,\eta)} \} = \frac{\pi {\cal N}_{\perp}}{L a \omega} \sum_{\tau,\tau'=\pm 1} \sum_{|q|,|k| \leq \pi/(2a)} \left| \lambda_{\tau,\tau'} (k,q)\right|^2 \delta(\omega - E_{\tau,\tau'}(k,q)) \end{equation} with the absorption energies between the respective Peierls subbands \begin{equation} E_{\tau,\tau'}(k,q) = U +\tau' E(k+q/2)-\tau E(k-q/2) \; , \label{Etautauprime} \end{equation} see eq.~(\ref{bandhldim}) and figure~\ref{HueckelHarrisLangedis}. The transition matrix elements are given by \begin{eqnarray} \left| \lambda_{\tau,\tau'} (k,q)\right|^2 &=& \frac{(ea)^2}{2L} \Biggl\{ g(q) \left| (1+\delta\eta) \epsilon(k)f_{\tau,\tau'} +\tau\tau' (\delta+\eta) \frac{\Delta(k)}{\delta}f_{\tau,-\tau'}\right|^2 \nonumber \\[3pt] && \phantom{ \frac{(ea)^2}{2L} \Biggl\{ } + g(q+\frac{\pi}{a}) \left| (\delta+\eta) \epsilon(k)f_{\tau,\tau'} +\tau\tau' (1+\delta\eta) \frac{\Delta(k)}{\delta}f_{\tau,-\tau'}\right|^2 \Biggr\} \label{thelambdassquared} \end{eqnarray} \end{mathletters}% with the help functions \begin{mathletters} \label{thefs} \begin{eqnarray} f_{+,+}(k,q) = f_{-,-}(k,q) &=& \alpha_{k+q/2}^{\phantom{*}}\alpha_{k-q/2}^* - \beta_{k+q/2}^{\phantom{*}}\beta_{k-q/2}^* \\[3pt] f_{+,-}(k,q) = f_{-,+}(k,q) &=& \alpha_{k+q/2}^{\phantom{*}}\beta_{k-q/2}^* + \beta_{k+q/2}^{\phantom{*}}\alpha_{k-q/2}^* \end{eqnarray} \end{mathletters}% where $\alpha_k$, $\beta_k$ are given in eq.~(\ref{alphabeta}). It can easily be checked that the case~$\delta=\eta=0$ is reproduced. For $\delta=1$, $\eta=0$ one recovers the result for the average optical conductivity of $L/2$ independent two-site systems since $E(k)=2t$, $\lambda_{+,-}(k,q)=0$, and $|\lambda_{+,+}(k,q)|^2 = (2tea)^2(g(q)+g(q+\pi/a))/(2L)$: \begin{equation} {\rm Re}\{\overline{\sigma(\omega >0, \delta=1,\eta=0)} \} = \frac{L}{2} \frac{{\cal N}_{\perp}}{La\omega} \pi \delta(\omega -U) \frac{(Wea)^2}{4} \end{equation} where we used $\int_{-\pi}^{\pi}dq/(2\pi)\, g(qa)=1$. For the direct calculation we have to recall that only the singlet of the four spin states contributes, and the hopping between the two sites is~$2t$. For general~$\delta$, $\eta$ it is necessary to evaluate the optical conductivity in eq.~(\ref{monstersigma}) numerically. An example is shown in figure~\ref{hl10}. It is seen that now there are two side-peaks in the optical absorption spectrum. The new peaks due to the Peierls distortion are weaker than the one at~$\omega=U$ and vanish for $\delta\to 1$. For large lattice distortions the dominant contribution to the side peaks comes from the small-$q$ transitions between different Peierls subbands. Their oscillator strength is maximum for $\omega=U\pm W\sqrt{\delta}$ which determines the position of the peaks for large $\delta$. The Peierls gap between the bands shows up in the optical spectrum. For small lattice distortions all $(k,q)$ contribute. The signature of the Peierls gap is smeared out and the position of the side peaks cannot be expressed in terms of a simple function of~$\delta$. \section{Optical absorption in the extended Harris-Lange model} \label{optabsHubbard} \subsection{Extended dimerized Harris-Lange model} Strongly isotropic, almost ideal one-dimensional systems like polymers or charge-transfer salts are not properly described by the Hubbard model of eq.~(\ref{Hubb-Model}) for two reasons: (i)~the Peierls distortion is not taken into account, and (ii)~the residual Coulomb interaction between charges beyond the Hubbard on-site interaction is neglected. The exponential decay of the Wannier wave functions naturally allows to limit the interactions to on-site and nearest-neighbor Hubbard terms. In the ``Zero Differential Overlap Approximation'' it is further {\em assumed\/} that only the direct Coulomb term has to be taken into account for the nearest-neighbor Coulomb interaction (Kivelson, Su, Schrieffer, and Heeger 1987); (Wu, Sun, and Nasu 1987); (Baeriswyl, Horsch, and Maki 1988); (Gammel and Campbell 1988); (Kivelson, Su, Schrieffer, and Heeger 1988); (Campbell {\em et al.} 1988); (Painelli and Girlando 1988); (Painelli and Girlando 1989); (Campbell, Gammel, and Loh 1990). Optical absorption spectra for the extended dimerized Hubbard model could only be calculated numerically for small system sizes. Within such an approach the Hamiltonian is explicitly diagonalized (Soos and Ramesesha 1984); (Tavan and Schulten 1986); (Guo, Mazumdar, Dixit, Kajzar, Jarka, Kawabe, and Peyghambarian 1993); (Guo, Guo, and Mazumdar 1994). Since the dimension of the Hilbert space increases exponentially (${\rm dim} \hat{H} = 4^L$) the numerical analysis is restricted to short chains ($L \leq 12$) due to the limited computer power. Therefore, it is natural to analytically investigate the (dimerized) Harris-Lange model with an additional nearest-neighbor interaction. We will show below that the optical spectrum can still be calculated for this model which is equivalent to the extended dimerized Hubbard model to order $t (t/U)^{-1}$, $t (t/U)^0$, and $t(V/U)^0$. The dimerized extended Harris-Lange model reads \begin{mathletters}\begin{eqnarray} \hat{H}_{\rm HL}^{\rm dim, ext} &=& \hat{T}_{\rm LHB}(\delta) + \hat{T}_{\rm UHB}(\delta) + U \hat{D} + V \hat{V} \\[6pt] \hat{V} &=& \sum_l (\hat{n}_l-1)(\hat{n}_{l+1}-1) \; . \end{eqnarray}\end{mathletters}% For half-filling the ground state of the extended dimerized Harris-Lange model is still $2^L$-fold spin degenerate because every site is singly occupied for $| V | < U/2$. The energy of these states is zero, $E_0=0$, irrespective of the dimerization value~$\delta$. The double occupancy and the hole in the excited states now experience a nearest-neighbor attraction while the spin sector remains unchanged. Thus we may immediately translate~$\hat{V}$ into our band picture as \begin{equation} \hat{V}^{\rm band} = - \frac{2}{L} \sum_{|q|\leq \pi/a}\cos(qa) \sum_{|k|,|p|\leq \pi/a} \hat{u}_{k+q}^+\hat{u}_{k}^{\phantom{+}} \hat{l}_{p}^{\phantom{+}} \hat{l}_{p-q}^+ \end{equation} which describes the scattering of a hole in the lower Hubbard band with a particle in the upper Hubbard band. Again, the projection operators can be disregarded for the optical absorption. \subsection{Equation of motion technique} In the presence of the nearest-neighbor interaction it becomes increasingly tedious to separately calculate the exact eigenenergies and oscillator strengths. We rather prefer to directly calculate the optical conductivity from an equation of motion approach. \subsubsection{Translational invariant case} Since we are interested in the real part of the optical conductivity we can concentrate on the particle current density, \begin{equation} \langle \hat{\jmath}_t \rangle = \frac{{\cal N}_{\perp}}{La} \left( \langle 0(t) | \hat{\jmath} | \Psi(t)\rangle + {\rm h.c.}\right) \end{equation} where $|0(t)\rangle$ and $|\Psi(t)\rangle$ are the time evolution of the ground state with and without the external perturbation and thus obey the corresponding Schr\"{o}dinger equations. We have already used the fact that we want to calculate the linear absorption. We write \begin{eqnarray} \langle 0(t) | \hat{\jmath} | \Psi(t)\rangle &=& \sum_{k,q} -iea\epsilon(k)\hat{x}_{q}^+ \langle 0(t) | \hat{l}_{k-q/2}^+\hat{u}_{k+q/2}^{\phantom{+}} | \Psi(t)\rangle \nonumber \\[3pt] & \equiv &\sum_{k,q} \lambda(k,q) j_{k;q}(t) \; . \end{eqnarray} Upon Fourier transformation we obtain \begin{equation} \langle \hat{\jmath}_{\omega} \rangle = \frac{{\cal N}_{\perp}}{La} \left( \sum_{k,q} \lambda(k,q) j_{k;q}(\omega) + \lambda^+(k,q) j_{k;q}^*(-\omega) \right) \; . \end{equation} Since we are interested in the optical conductivity for positive frequencies (optical absorption) we may disregard the second term which contributes to~$\omega < 0$. The equation of motion for~$j_{k;q}(t)$ becomes \begin{mathletters} \begin{eqnarray} i \frac{\partial j_{k;q}(t) }{\partial t} &=& \langle 0(t) | \left[ \hat{l}_{k-q/2}^+\hat{u}_{k+q/2}^{\phantom{+}}, \hat{H}_{\rm HL}^{\rm band} \right]_{-} | \Psi(t)\rangle - \frac{{\cal A}(t)}{c} \langle 0(t) | \hat{l}_{k-q/2}^+\hat{u}_{k+q/2}^{\phantom{+}} \hat{\jmath} | 0(t)\rangle \\[6pt] \omega j_{k;q}(\omega) &=& E(k,q) j_{k;q}(\omega) - 2V\left( \cos(ka) j_q^c(\omega) +\sin(ka)j_q^s(\omega)\right) - \lambda^+(k,q) \frac{{\cal A}(\omega)}{c} \end{eqnarray} \end{mathletters}% where we kept the expansion linear in the external perturbation, and performed the Fourier transformation. Furthermore, we introduced the abbreviations \begin{equation} j_q^{c,s}(\omega) = \frac{1}{L} \sum_k \left( {\cos (ka) \atop \sin (ka)}\right) j_{k;q}(\omega) \; . \end{equation} For our calculations we only need~$j_q^c(\omega)$ since our current operator preserves parity. The particle current density for positive frequencies becomes \begin{equation} \langle \hat{\jmath}_{\omega>0}\rangle = \frac{{\cal N}_{\perp}}{a}(2tiea) \sum_q \hat{x}_q^+ j_{q}^c(\omega) \end{equation} which is proportional to the external field. We introduce the function \begin{equation} F(q) = \frac{2}{L} \sum_{|k|\leq \pi/a} \frac{(\cos ka)^2}{\omega - E(k,q)} \end{equation} which allows us to finally express the optical conductivity as \begin{equation} {\rm Re} \{ \overline{\sigma(\omega>0,V)} \} = - \frac{(Wea)^2{\cal N}_{\perp}}{16 a\omega} \frac{1}{L} \sum_{|q|\leq \pi/a} g(q) {\rm Im}\left\{ \frac{F(q)}{1+VF(q)} \right\} \; . \label{ResigmaHLV} \end{equation} The result will be discussed in the next subsection. \subsubsection{Extended dimerized Harris-Lange model} The same procedure can be applied to the dimerized case where it is best to start from the diagonalized Hamiltonian in the form of eq.~(\ref{effHL}), and the current operator in the form of eq.~(\ref{jinterdimHL}). The calculations are outlined in appendix~\ref{appc}. We introduce the three functions $F_{1,2,3}$ as \begin{mathletters} \label{capitalF} \begin{eqnarray} F_{1} (q) &=& \frac{2}{L} \sum_{|k| \leq \pi/(2a)} \cos^2(ka) \Biggl[ |f_{+,+}|^2 \left( \frac{1}{\omega-E_{-,-}} +\frac{1}{\omega-E_{+,+}} \right) \nonumber \\[3pt] && \phantom{\frac{2}{L} \sum_{|k| \leq \pi/(2a)} \cos^2(ka) \Biggl[ } +|f_{+,-}|^2 \left(\frac{1}{\omega-E_{-,+}}+\frac{1}{\omega-E_{+,-}}\right) \Biggr] \\[6pt] F_{2} (q) &=& \frac{2}{L} \sum_{|k| \leq \pi/(2a)} \sin^2(ka) \Biggl[ |f_{+,-}|^2 \left( \frac{1}{\omega-E_{-,-}} +\frac{1}{\omega-E_{+,+}} \right) \nonumber \\[3pt] && \phantom{\frac{2}{L} \sum_{|k| \leq \pi/(2a)} \sin^2(ka) \Biggl[ } +|f_{+,+}|^2 \left(\frac{1}{\omega-E_{-,+}}+\frac{1}{\omega-E_{+,-}}\right) \Biggr] \\[6pt] F_{3} (q) &=& \frac{2}{L} \sum_{|k| \leq \pi/(2a)} \cos(ka)\sin(ka) \Biggl\{ f_{+,+}^{\phantom{*}} f_{+,-}^* \left[ \frac{1}{\omega-E_{+,+}}+ \frac{1}{\omega-E_{-,-}} \right] \nonumber \\[3pt] && \phantom{\frac{2}{L} \sum_{|k| \leq \pi/(2a)} \cos(ka)\sin(ka) \biggl\{ } - f_{+,+}^* f_{+,-}^{\phantom{*}} \left[ \frac{1}{\omega-E_{-,+}}+ \frac{1}{\omega-E_{+,-}} \right]\Biggr\} \end{eqnarray} \end{mathletters}% where $f_{\tau,\tau'}\equiv f_{\tau,\tau'}(k,q)$ and $E_{\tau,\tau'}=E_{\tau,\tau'}(k,q)$ were introduced in eq.~(\ref{thefs}) and eq.~(\ref{Etautauprime}). Furthermore, we abbreviate~$A_j=(1+\delta\eta)F_j-(\eta+\delta)F_3$, $B_j=(\delta+\eta)F_j-(1+\delta\eta)F_3$ ($j=1,2$), $C_1=(1+\delta\eta)^2 F_1+(\delta+\eta)^2 F_2 -2(1+\delta\eta)(\delta+\eta)F_3$, and $C_2=(1+\delta\eta)^2 F_2+(\delta+\eta)^2 F_1 -2(1+\delta\eta)(\delta+\eta)F_3$. The real part of the average optical conductivity can then be expressed as \begin{eqnarray} {\rm Re}\{\overline{\sigma(\omega >0, V, \delta,\eta)} \} &=& {\rm Re}\{\overline{\sigma(\omega >0,\delta,\eta)} \} \nonumber \\[6pt] && + \frac{V{\cal N}_{\perp}(Wea)^2}{16 a\omega L} {\rm Im}\Biggl\{ \sum_{|q| \leq \pi/(2a)} \frac{1}{(1+VF_1)(1+VF_2)-(VF_3)^2} \label{thefinalresultHL} \\[6pt] && \phantom{+ \frac{V{\cal N}_{\perp}(Wea)^2}{16 a\omega L} {\rm Im}\Biggl\{ } \biggl\{ g(q) \left[ A_1^2 +B_2^2 +V(F_1F_2-F_3^2) C_1 \right] \nonumber \\ && \phantom{+ \frac{V{\cal N}_{\perp}(Wea)^2}{16 a\omega L} {\rm Im}\Biggl\{ \biggl\{ } +g(q+\frac{\pi}{a}) \left[ A_2^2 +B_1^2 +V(F_1F_2-F_3^2) C_2 \right] \biggr\} \Biggr\} \; . \nonumber \end{eqnarray} The result for~$V=0$ is given in eq.~(\ref{monstersigma}). In the following we will discuss the results for the average optical absorption in the presence of a nearest-neighbor interaction. \subsection{Optical absorption} \subsubsection{Translational invariant case} The help function~$F(q)$ can be calculated analytically with the help of eqs.~(2.267,1), (2.266), and~(2.261) of (Gradshteyhn and Ryzhik 1980). The result is \begin{equation} F(q) \! = \! \frac{2}{[4t\sin(qa/2)]^2} \! \left\{ \! \begin{array}{lcl} \omega-U -\sqrt{(\omega-U)^2- \left(4t\sin(qa/2)\right)^2\, } & {\rm for} & |\omega-U| \geq |4t\sin(qa/2)| \\[6pt] -i \sqrt{ \left(4t\sin(qa/2)\right)^2 - (\omega-U)^2\, } & {\rm for} & |\omega-U| < |4t\sin(qa/2)| \end{array} \right. \, . \end{equation} The result for~$V=0$, eq.~(\ref{what}), follows after the substitution of~$x=\sin(qa/2)$ into equation~(\ref{ResigmaHLV}). The total optical absorption is shown in figure~\ref{hl01}. For arbitrarily small~$V>0$ there is a bound exciton which is the standard situation for one-dimensional short-range attractive potentials between a positive (hole) and negative charge (double occupancy). This is evident from the form of~$F(q)$ which allows for excitons with momenta~$q$ if $|\omega -U| \geq |4t\sin(qa/2)|$ which is fulfilled for~$q=0$ for all~$V>0$. When the attraction between the two opposite charges is strong, the full exciton band with width~$W_{\rm exc}=4t^2/V$ is formed. This can be seen from the zeros of the denominator in eq.~(\ref{ResigmaHLV}) in the region~$|\omega-(U-V-W_{\rm exc}/2)|\leq W_{\rm exc}/2$. One finds from $1+VF(q)=0$ that \begin{eqnarray} \omega= U-V-\frac{W_{\rm exc}}{2} + \frac{W_{\rm exc}}{2} \cos qa \; . \end{eqnarray} This is precisely the dispersion relation for bound pairs in one dimension with nearest-neighbor hopping of strength~$t_{\rm exc}=t (t/V)$: at large~$V$ the excitons are essentially nearest-neighbor pairs of opposite charges which {\em coherently\/} move with the hopping amplitude~$t_{\rm exc}$. Note that this motion requires an intermediate (``virtual'') configuration where the two charges are not nearest neighbors. Consequently, the hopping integral of the individual constituents,~$t$, is reduced by the factor~$t/V$ for the motion of the pair. The full band becomes apparent when $W_{\rm exc}+V > W$ or $V > W/2$, see figure~\ref{hl01}. Recall that the spin sector provides {\em any\/} momentum to the charge sector. The momentum transfer, however, is modulated by the function~$g(q)$ which is maximum at $q=\pi/a$ and reflects the fact that states with antiferromagnetic spin correlations are best suited for optical absorptions since they contain many neighboring singlet pairs. Hence, the $q=\pi/a$-exciton dominates over the $q=0$-exciton for $V>W/2$. It is amusing to see that the optical absorption of a Peierls insulator and a Mott-insulator (extended Harris-Lange model: $U \gg W$, $V> W/2$, $J=0$) can look very similar, compare figure~2 of~I and figure~\ref{hl01}. This has already been noted long time ago by Simpson (Simpson 1951); (Simpson 1955); (Salem 1966); (Fave 1992) who explained the optical absorption spectra of short polyenes in the above exciton model. It is seen that Simpson's model is naturally included in our strong-correlation approach. For real polymers, however, Simpson's original approach is not satisfactory. A fully developed exciton band only exists in the presence of an incoherent spin background. Now that even the spin-Peierls effect is excluded one can by no means explain the Peierls distortion of the lattice as an electronic effect. This does not exclude other, e.g., extrinsic, explanations for a lattice distortion. \subsubsection{Extended dimerized Harris-Lange model} The full spectrum has to be determined numerically. An example for various values of~$V/t$ is shown in figure~\ref{hl11}. For large~$V/t$ we have a fully developed exciton band which is itself Peierls-split into two branches. Thus one obtains four van-Hove singularities in the optical absorption spectrum. The Peierls gap is given by~$\Delta_{\rm exc}^{\rm P}= \delta W_{\rm exc}= 4t^2\delta/V$. Even for~$V=W$ it is smeared out since $V/t$ is not too large yet and the phenomenological damping~$\gamma$ is already of the order of the gap. For small~$V/t$ we obtain the signature of the $q=0$ exciton for~$\delta=0$, $V>0$. For intermediate~$V$ this peak develops into a van-Hove singularity of the upper exciton subband. The signatures of the second van-Hove singularity of the upper band are clearly visible for~$V=W/2$. The peaks of the Peierls subbands for~$V=0$, $\delta \neq 0$ are both red-shifted. The peak at lower energy increases in intensity and finally forms the lower exciton subband while the peak at higher energy quickly looses its oscillator strength. \section{Summary and Outlook} In this paper we addressed the optical absorption of the half-filled Harris-Lange model which is equivalent to the Hubbard model at strong correlations and temperatures large compared to the spin energy scale. It is extremely difficult to analytically calculate optical properties of interacting electrons in one dimension. For strong coupling when the Hubbard interaction is large compared to the band-width matters considerably simplified since the energy scales for the charge and spin excitations are well separated. We were able to derive an {\em exactly\/} equivalent band structure picture for the charge degrees of freedom and found that the upper and lower Hubbard band are actually {\em parallel\/} bands with the band structure of free Fermions. We have taken special care of the spin background which can act as a momentum reservoir for the charge system. Since we can exactly integrate out the spin degrees of freedom for the Harris-Lange model we were able to solve the problem even in the presence of a lattice dimerization and a nearest-neighbor interaction between the electrons. For a vanishing nearest-neighbor interaction we found a prominent absorption peak at~$\omega=U$ and additional side peaks in the absorption bands $|\omega-U| \leq W$ in the presence of a lattice distortion~$\delta$. When a further nearest-neighbor interaction between the charges was included, we found the formation of Simpson's exciton band of band-width $W_{\rm exc.}=4t^2/V$ for $V>W/2$ which is eventually Peierls-split. As usual the excitons draw almost all oscillator strength from the band. It should be clear that the Harris-Lange model with its highly degenerate ground state is not a suitable model for the study of real materials. The results presented here are relevant to systems for which the temperature is much larger than the spin exchange energy. Real experiments are not carried out in this ``hot-spin'' regime but at much lower temperatures for which the system is in an unique ground state with antiferromagnetic correlations. Unfortunately, this problem cannot be solved analytically. In the third and last paper of this series (Gebhard {\em et al.} III 1996) we will employ the analogy to an ordinary semiconductor (electrons and holes in a phonon bath) to design a ``no-recoil'' approximation for the chargeons in a spinon bath. It will allows us to determine the coherent absorption features of the Hubbard model at large~$U/t$. \section*{Acknowledgments} We thank H.~B\"{a}\ss ler, A.~Horv\'{a}th, M.~Lindberg, S.~Mazumdar, M.~Schott, and G.~Weiser for useful discussions. The project was supported in part by the Sonderforschungsbereich~383 ``Unordnung in Festk\"{o}rpern auf mesoskopischen Skalen'' of the Deutsche Forschungsgemeinschaft. \newpage \begin{appendix} \section{The Harris-Lange model} \label{appeigen} \subsection{Sum rules} \label{appsumrule} We briefly account for the sum rules. We have \begin{equation} \int_{0}^{\infty} d\omega\ {\rm Im}\{\chi(\omega)\} = \pi \frac{{\cal N}_{\perp}}{La} \sum_n \left| \langle 0 | \hat{\jmath}^2|n\rangle\right|^2 = \pi \frac{{\cal N}_{\perp}}{La} \langle 0 | \hat{\jmath}^2|0\rangle \; . \end{equation} It is a standard exercise to show that \begin{equation} \langle 0 | \hat{\jmath}^2 | 0 \rangle = (2tea)^2 \sum_l \left(1+(-1)^l\delta\right)^2\left(1+(-1)^l\eta\right)^2 \langle 0 | \left( \frac{1}{4} -\hat{\rm\bf S}_l\hat{\rm\bf S}_{l+1}\right) | 0 \rangle \; . \end{equation} We define the positive quantities \begin{equation} C_S^{\rm even, odd}= \frac{1}{L} \sum_l \frac{1\pm (-1)^l}{2} \langle 0 | \left( \frac{1}{4} -\hat{\rm\bf S}_l\hat{\rm\bf S}_{l+1}\right) | 0 \rangle \label{CSevenCSodd} \end{equation} and may then write \begin{equation} \int_{0}^{\infty} d\omega\ {\rm Im}\{\chi(\omega)\} = \pi {\cal N}_{\perp}a (2te)^2 \left[ \left(1+\delta\right)^2\left(1+\eta\right)^2 C_S^{\rm even} + \left(1-\delta\right)^2\left(1-\eta\right)^2 C_S^{\rm odd} \right] \; . \end{equation} If we average over all possible states~$|0\rangle$ we obtain $C_S^{\rm even, odd}=1/8$ since only the singlet configuration contributes. Hence, \begin{equation} \int_{0}^{\infty} d\omega \overline{{\rm Im}\{\chi(\omega)\}} = \pi {\cal N}_{\perp}a (te)^2 \left[(1+\delta\eta)^2 +(\delta+\eta)^2\right] \; . \end{equation} The area under the curves for~$\sigma_{\rm red}(\omega)$, eq.~(\ref{sigmared}), are thus given by \begin{mathletters} \begin{eqnarray} \int_0^{\infty} \frac{d\omega}{W} \sigma_{\rm red}(\omega) &=& \frac{\pi}{4} \left[ \left(1+\delta\right)^2\left(1+\eta\right)^2 C_S^{\rm even} + \left(1-\delta\right)^2\left(1-\eta\right)^2 C_S^{\rm odd} \right] \label{sumrulesigmared} \\[6pt] \int_0^{\infty} \frac{d\omega}{W} \overline{\sigma_{\rm red}(\omega)} &=& \frac{\pi}{16} \left[(1+\delta\eta)^2 +(\delta+\eta)^2\right] \; . \end{eqnarray} \end{mathletters}% \subsection{Momentum of eigenstates} \label{appmomentum} First we will assume that the number of sites and the number of particles is even. $N_C$ and~$N_S$ will thus also be even. Let~$\hat{{\cal T}}$ be the translation operator by one site, i.~e., $\hat{{\cal T}} \hat{c}_{l,\sigma}\hat{{\cal T}}^{-1}= \hat{c}_{l+1,\sigma}$. We have to show that \begin{equation} \hat{{\cal T}} |\Psi\rangle = e^{-iPa}|\Psi\rangle \end{equation} holds. We have to distinguish two cases: (i)~a spin is at site~$L$ in $|\Psi\rangle$ or (ii)~a charge is at site~$L$ in $|\Psi\rangle$. \paragraph{case~(i):} the operator~$\hat{{\cal T}}$ shifts the spin from site~$L$ to the first site. This results in a phase factor~$(-1)$ since one has to commute the Fermion operator~$(N-1)$-times to obtain the proper order in~$|\Psi\rangle$. Furthermore, the states in~$|\Psi\rangle$ have, relative to those in $\hat{{\cal T}}|\Psi\rangle$, \begin{enumerate} \item shifted the spin sequence by one unit. This results in a phase factor $\exp(ik_Sa)$; \item an additional factor $(-1)$ for each doubly occupied site. This gives a phase factor~$\exp(i\pi N_d)$; \item a Slater determinant in which each site index is shifted by one. This results in a phase factor $\exp(i\sum_{j=1}^{N_C} (k_j+\Phi_{CS})a)$. \end{enumerate} In sum we obtain \begin{eqnarray} (-1) &=& e^{-iPa} e^{ik_Sa} e^{i\pi N_d} e^{i\sum_{j=1}^{N_C} (k_j+\Phi_{CS})a} \nonumber \\[6pt] P&=& k_S +(\pi/a) (N_d-1)+ \sum_{j=1}^{N_C} (k_j+\Phi_{CS}) \quad {\rm mod\ } 2\pi/a \; . \end{eqnarray} \paragraph{case~(ii):} the operator~$\hat{{\cal T}}$ shifts the charge from site~$L$ to the first site. The states in~$|\Psi\rangle$ have, relative to those in $\hat{{\cal T}}|\Psi\rangle$, \begin{enumerate} \item shifted the charge sequence by one unit. This results in a phase factor $\exp(ik_Ca)$; \item an additional factor $(-1)$ for each doubly occupied site. This gives a phase factor~$\exp(i\pi N_d)$; \item a Slater determinant in which \begin{enumerate} \item each site index is shifted by one. This results in a phase factor $\exp(i\sum_{j=1}^{N_C} (k_j+\Phi_{CS})a)$; \item the last row and the first row are interchanged. This gives an additional factor $(-1)^{N_C-1}=-1$; \item the first row is $(1,\ldots 1)$ instead of $\left(\exp(i(k_1+\Phi_{CS})La),\ldots\exp(i(k_{N_C}+\Phi_{CS})La)\right)$ $= (1,\ldots 1)\exp(i(k_C-k_C)a)$. This gives an additional phase factor~$\exp(-i(k_C-k_S)a)$. \end{enumerate} \end{enumerate} In sum we obtain \begin{eqnarray} 1 &=& e^{-iPa} e^{ik_C a} e^{i\pi N_d} e^{i\sum_{j=1}^{N_C} (k_j+\Phi_{CS})a} (-1) e^{-i(k_C-k_S)a} \nonumber \\[6pt] P&=& k_S +(\pi/a) (N_d-1)+ \sum_{j=1}^{N_C} (k_j+\Phi_{CS}) \quad {\rm mod\ } 2\pi/a \end{eqnarray} as before. If the number of particles~$N$ is odd, the momentum is shifted by another factor of~$\pi/a$, if one repeats the above arguments. This proves that the total momentum of the state~$|\Psi\rangle$ is indeed given by eq.~(\ref{eigenmomHL}). \subsection{Energy of eigenstates} \label{appenerg} We want to prove that \begin{equation} \hat{H}_{\rm HL} |\Psi\rangle = E|\Psi\rangle \quad ; \quad E = \sum_{j=1}^{N_C} \epsilon(k_j+\Phi_{CS}) +U N_d \; . \label{showit} \end{equation} Again we restrict ourselves to even~$N$. The bulk terms are simple since there is no hopping across the boundary. A hopping process of a double occupancy and a hole are equivalent: \begin{mathletters} \begin{eqnarray} (-1)^l | \ldots \bullet_l \sigma_{l+1} \ldots\rangle &\mapsto& - (-1)^l | \ldots \sigma_{l} \bullet_{l+1} \ldots\rangle = (-1)^{l+1} | \ldots \sigma_{l} \bullet_{l+1} \ldots\rangle \\[3pt] | \ldots \circ_l \sigma_{l+1} \ldots\rangle &\mapsto& | \ldots \sigma_{l} \circ_{l+1} \ldots\rangle \; . \end{eqnarray} \end{mathletters}% The extra minus sign which appears when a double occupancy moves has been taken care of in the wave function by the phase factor~$(-1)^l$ for a double occupancy at site~$l$. Now that there is no difference in the motion of double occupied sites and holes they dynamically behave as spinless Fermions. Since the Slater determinant is the proper phase factor for non-interacting Fermions eq.~(\ref{showit}) holds for the bulk terms. This also shows that only the Harris-Lange model with hopping amplitudes~$|t_{\rm LHB}|=|t_{\rm UHB}|$ can be solved. Another integrable but trivial case is~$t_{\rm LHB}=0$, $t_{\rm UHB}\neq 0$ and vice versa. We now address the boundary terms. A typical configuration for which transport across the boundary is possible is $|S_1, \ldots C_L\rangle$. The phase of the configuration is given by a Slater determinant in which the last row has the entry $\left(\exp(i(k_1+\Phi_{CS})La),\ldots\exp(i(k_{N_C}+\Phi_{CS})La)\right)$ $= (1,\ldots 1)\exp(i(k_C-k_S)La)$. The action of~$\hat{H}_{\rm HL}$ moves the charge from site~$L$ to the first position by which an extra minus sign occurs since the electron operator for the spin had to be commuted with~$(N-1)$ other electron operators. These phase factors have to be compared to the corresponding configuration in~$E|\Psi\rangle$. Relative to the configuration in~$\hat{H}_{\rm HL}|\Psi\rangle$ it has \begin{enumerate} \item shifted the spin sequence by minus one unit. This results in a phase factor~$\exp(-ik_Sa)$; \item shifted the charge sequence by one unit. This results in a phase factor~$\exp(ik_Ca)$; \item a Slater determinant which has $\left(\exp(i(k_1+\Phi_{CS})a),\ldots\exp(i(k_{N_C}+\Phi_{CS})a)\right)$ in the first row. \end{enumerate} The boundary terms should give the same result as the bulk terms. This leads to the condition \begin{equation} - e^{i(k_C-k_S)La} = e^{-ik_Sa} e^{ik_C a} (-1)^{N_C-1} \end{equation} which is obviously fulfilled. The proof for odd~$N$ is analogous, and eq.~(\ref{showit}) holds for all~$N$. \subsection{Electrical dipole operator for the Harris-Lange model} \label{appdipoleHL} We can derive the electrical dipole operator for the Harris-Lange model from its definition in equation~(A.22) of~I. We use the Hamilton operator in the band picture interpretation, eq.~(\ref{effHL}), and the corresponding current operator, eq.~(\ref{jhlbandpicture}). Since~$\hat{x}_q$ can be replaced by its average value~$\sqrt{g(q)/(2L)}$ one easily sees that the dipole operator becomes \begin{mathletters} \begin{eqnarray} \hat{P}_{\rm inter}^{\rm HL}&=&\sum_{|k|,|q|\leq \pi/a} \mu_{\rm inter}^{\rm HL}(k,q) \left( \hat{u}_{k+q/2}^+\hat{l}_{k-q/2}^{\phantom{+}} + \hat{l}_{k-q/2}^+ \hat{u}_{k+q/2}^{\phantom{+}} \right) \\[3pt] \mu_{\rm inter}^{\rm HL}(k,q)&=& i \frac{\lambda(k,q)}{E(k,q)}= ea \sqrt{\frac{g(q)}{2L}} \frac{\epsilon(k)}{E(k,q)} \end{eqnarray} \end{mathletters}% with~$E(k,q)=U+\epsilon(k+q/2)-\epsilon(k-q/2)$. One sees that the dipole matrix element is of the order~$t/U$ as it should be for interband transitions. Furthermore, for small momentum transfer we obtain \begin{equation} \mu_{\rm inter}^{\rm HL}(k;q\to 0) \sim \epsilon(k) \; . \end{equation} This is the correct form since we create a neighboring hole and double occupancy which corresponds to an electric dipole between nearest neighbors. The procedure is readily generalized to the dimerized Harris-Lange model. The interband current operator in terms of the Fermion operators for the four Peierls subbands is given in eq.~(\ref{jinterdimHL}), and the diagonalized Hamiltonian can be found in eq.~(\ref{bandhldim}). One readily finds \begin{mathletters} \begin{eqnarray} \hat{P}_{\rm inter}^{\rm dim.\ HL}&=& \sum_{\tau,\tau'=\pm 1} \sum_{|k|,|q|\leq \pi/(2a)} \mu_{\rm inter;\tau,\tau'}^{\rm dim.\ HL}(k,q) \left( \hat{u}_{k+q/2,\tau'}^+\hat{l}_{k-q/2,\tau}^{\phantom{+}} + \hat{l}_{k-q/2,\tau}^+\hat{u}_{k+q/2,\tau'}^{\phantom{+}} \right) \\[3pt] \mu_{\rm inter;\tau,\tau'}^{\rm HL}(k,q)&=& i \frac{\lambda_{\tau,\tau'}(k,q)}{E_{\tau,\tau'}(k,q)} \end{eqnarray} \end{mathletters}% with $\lambda_{\tau,\tau'}(k,q)$ as the root of eq.~(\ref{thelambdassquared}), and $E_{\tau,\tau'}(k,q)=U+\tau' E(k+q/2)-\tau E(k-q/2)$, see eq.~(\ref{Etautauprime}). The dipole matrix elements again simplify for small~$q$. Note that one obtains both contributions from $q\to 0$ and $q\to \pi/a$. After some calculations one obtains \begin{mathletters} \label{lambdatautauprimeqzero} \begin{eqnarray} \left|\lambda_{+,+}(k;q=0)\right|^2 &=& \frac{1}{2L} \Biggl\{ \frac{1}{3} \left[ea\left( E(k)+\delta\eta\frac{(2t)^2}{E(k)}\right) \right]^2 \nonumber \\[6pt] && \phantom{ \frac{1}{4L} \biggl[ } + 3 \left[ea \left(\eta E(k)-\delta\frac{(2t)^2}{E(k)}\right) \right]^2 \Biggr\} \\[6pt] \left|\lambda_{+,-}(k;q= 0)\right|^2 &=& \frac{1}{2L} \left( ea \frac{\epsilon(k)\Delta(k)(1-\delta^2)}{\delta E(k)}\right)^2 \left( \frac{\eta^2}{3} +3\right) \; . \end{eqnarray} \end{mathletters}% Note that the dipole matrix elements $\left|\lambda_{+,-}(k;q=0)\right|^2$ contain the contributions from~$q= \pi/a$ for $\delta=\eta=0$. Eqs.~(\ref{lambdatautauprimeqzero}) have to be compared to the corresponding expressions for the Peierls chain. It is seen that the expressions display some similarities but they show subtle differences. Even for~$q=0$ the expressions~(\ref{lambdatautauprimeqzero}) could not have been guessed. The corresponding dipole matrix elements become \begin{mathletters} \label{mutautauprimeqzero} \begin{eqnarray} \left|\mu_{+,+}(k;q\to 0)\right|^2 &=& \frac{1}{U^2} \left|\lambda_{+,+}(k;q\to 0)\right|^2\\[6pt] \left|\mu_{+,-}(k;q\to 0)\right|^2 &=& \frac{1}{(U-2E(k))^2} \left|\lambda_{+,-}(k;q\to 0)\right|^2 \\[6pt] \left|\mu_{-,+}(k;q\to 0)\right|^2 &=& \frac{1}{(U+2E(k))^2} \left|\lambda_{+,-}(k;q\to 0)\right|^2 \; . \end{eqnarray} \end{mathletters}% The dipole matrix elements between the same Peierls subbands are always strong, irrespective of~$k$ or $\delta$. However, the dipole matrix elements for transitions between different subbands are small for strong dimerization. Furthermore, they are small in the vicinity of the center and the edge of the reduced Brillouin zone. \subsection{Spin average in the Harris-Lange model} \label{spiav} We need to calculate \begin{equation} \langle \hat{x}_q^+\hat{x}_{q'}\rangle = \frac{1}{2^L}\sum_{|0\rangle} \hat{x}_q^+\hat{x}_{q'} \end{equation} where {\arraycolsep=0pt\begin{eqnarray} \hat{x}_q^{+}(\delta,\eta)\hat{x}_{q'}^{\phantom{+}}(\delta,\eta) &=& \sum_{S_1^{\prime},\ldots S_{L-2}^{\prime}} \frac{1}{L^2} \sum_{l,r} e^{i(ql-q'r)a} \bigl(1+\eta\delta +(-1)^l(\delta+\eta)\bigr) \bigl(1+\eta\delta +(-1)^r(\delta+\eta)\bigr) \nonumber \\[6pt] && \phantom{\sum_{S_1^{\prime},\ldots S_{L-2}^{\prime}} \frac{1}{L^2} \sum_{l,r} e^{i(ql-q'r)} } \langle 0 | S_1^{\prime},\ldots S_{l-1}^{\prime}, \left( \uparrow_{l}\downarrow_{l+1}-\downarrow_{l}\uparrow_{l+1}\right), S_{l}^{\prime},\ldots S_{L-2}^{\prime}\rangle \nonumber \\[6pt] && \phantom{\sum_{S_1^{\prime},\ldots S_{L-2}^{\prime}} \frac{1}{L^2} \sum_{l,r} e^{i(ql-q'r)} } \langle S_{L-2}^{\prime},\ldots S_{r}^{\prime}, \left( \downarrow_{r+1} \uparrow_{r}-\uparrow_{r+1}\downarrow_{r}\right), S_{r-1}^{\prime},\ldots S_{1}^{\prime} | 0 \rangle \; . \nonumber \\ && \label{xqxqprimeapp} \end{eqnarray}}% Since the set of spin states $|0\rangle$ is complete we may exactly trace it out and are left with the calculation of the spin matrix elements \begin{eqnarray} &&M(l,r)=\frac{1}{2^L}\sum_{S_1^{\prime},\ldots S_{L-2}^{\prime}} \\[3pt] && \langle S_{L-2}^{\prime},\ldots S_{r}^{\prime}, \left( \downarrow_{r+1} \uparrow_{r}-\uparrow_{r+1}\downarrow_{r}\right), S_{r-1}^{\prime},\ldots S_{1}^{\prime} | S_1^{\prime},\ldots S_{l-1}^{\prime}, \left( \uparrow_{l}\downarrow_{l+1}-\downarrow_{l}\uparrow_{l+1}\right), S_{l}^{\prime},\ldots S_{L-2}^{\prime} \rangle \; . \nonumber \end{eqnarray}% The value of all spins between the sites~$l$ and~$r$ is fixed by the singlet operators at $(l, l+1)$ and $(r, r+1)$. We find \begin{equation} M(l,r) = 2(-1)^{r-l} 2^{-|r-l|-2} \end{equation} which shows that the correlation function for finding two singlet pairs at distance~$n=|r-l|$ exponentially decays with correlation length~$\xi_{\rm S}=1/\ln(2)$. Since $M(l,r)$ only depends on the distance between the two sites we may carry out one of the lattice sums in equation~(\ref{xqxqprimeapp}). This gives \begin{eqnarray} \langle \hat{x}_q^+\hat{x}_{q'}\rangle = \frac{1}{2L}\biggl\{ && \delta_{q,q'} \sum_{n=-L/2}^{L/2-1} e^{inqa}2^{-|n|}(-1)^n \left[(1+\delta\eta)^2+(-1)^n(\delta+\eta)^2\right] \\[3pt] \nonumber && + \delta_{q,q'+\frac{\pi}{a}} \sum_{n=-L/2}^{L/2-1} e^{inqa}2^{-|n|} (1+\delta\eta)(\delta+\eta)(1+(-1)^n)\biggr\} \; . \end{eqnarray} The sum over~$n$ is readily taken and gives the final result \begin{mathletters} \begin{eqnarray} \langle \hat{x}_q^+\hat{x}_{q'}\rangle &=& \frac{1}{2L} \biggl\{ \delta_{q,q'} \left[(1+\eta\delta)^2 g(q)+(\delta+\eta)^2 g(q+\frac{\pi}{a})\right] \nonumber \\[3pt] && \phantom{\frac{1}{2L} \biggl\{ } + \delta_{q,q'+\frac{\pi}{a}} (1+\eta\delta)(\delta+\eta) \left[g(q)+ g(q+\frac{\pi}{a})\right] \biggr\} \end{eqnarray} with the help function \begin{equation} g(q) = \frac{3}{5+4\cos(qa)} \; . \end{equation}\end{mathletters}% In particular, for the translational invariant case we find \begin{equation} \langle \hat{x}_q^+\hat{x}_{q'}\rangle = \delta_{q,q'}\frac{g(q)}{2L} \; . \end{equation} This shows that $\hat{x}_q$ can be replaced by $\sqrt{g(q)/(2L)}$ in the translational invariant case. \section{Equation of motion technique for the extended dimerized Harris-Lange model} \label{appc} Here we briefly outline the calculations for the most general case~$V\neq 0$, $\delta\neq 0$. We use the diagonalized band picture Hamiltonian in the form of eq.~(\ref{effHL}), and the band picture current operator in the form of eq.~(\ref{jinterdimHL}). This has the advantage that the equation of motion can directly be inverted and the contribution for~$V=0$ can immediately be separated. The equations of motions give for the four currents \begin{equation} j_{\tau,\tau'}(k,q;\omega) = -\frac{ ({\cal A}(\omega)/c)\lambda_{\tau,\tau'}^+ + 2V \left( \cos(ka) X_{\tau,\tau'}^c +\sin(ka) X_{\tau,\tau'}^s \right) } {\omega- E_{\tau,\tau'}} \label{C1} \end{equation} where~$E_{\tau,\tau'}\equiv E_{\tau,\tau'}(k,q)= U+\tau' E(k+q/2)-\tau E(k-q/2)$, $\lambda_{\tau,\tau'}\equiv\lambda_{\tau,\tau'}(k,q)$, and $X_{\tau,\tau'}^{c,s}\equiv X_{\tau,\tau'}^{c,s}(k,q)$. The difficult terms~$X^{c,s}\equiv X_{+,+}^{c,s}=-X_{-,-}^{c,s}$, $Y^{c,s}\equiv X_{+,-}^{c,s}=X_{-,+}^{c,s}$, come from the nearest-neighbor interaction which mixes excited pairs in the different Peierls subbands. With the help of eq.~(\ref{C1}), $\lambda_{+,+}=-\lambda_{-,-}$, and $\lambda_{+,-}=\lambda_{-,+}$ we can immediately write \begin{eqnarray} \langle \hat{\jmath}_{\omega >0}\rangle(V) &=& \langle \hat{\jmath}_{\omega >0}\rangle(V=0) \nonumber \\[6pt] && - \frac{2V{\cal N}_{\perp}}{La} \sum_{{|q| \leq \pi/(2a)} \atop {|k| \leq \pi/(2a)} } \left( {\cos(ka) \atop \sin(ka)} \right) \Biggl\{ \lambda_{+,+}X^{c,s} \left( \frac{1}{\omega-E_{+,+}} + \frac{1}{\omega-E_{-,-}} \right) \\[6pt] \label{C2} && \phantom{ - \frac{2V{\cal N}_{\perp}}{La} \sum_{|q| \leq \pi/(2a)} \left( {\cos(ka) \atop \sin(ka)} \right) \Biggl\{ } + \lambda_{+,-}Y^{c,s} \left( \frac{1}{\omega-E_{+,-}} + \frac{1}{\omega-E_{-,+}} \right) \Biggr\} \; . \nonumber \end{eqnarray} To determine~$X^{c,s}$ and $Y^{c,s}$ we have to evaluate $\hat{V} \hat{u}_{k+q/2,\tau'}^+ \hat{l}_{k-q/2,\tau}^{\phantom{+}} |0\rangle$. Since $\hat{V}$ is simple in terms of the original operators~$\hat{u}_{k}$, $\hat{l}_{k}$ we first have to apply the inverse transformation of eq.~(\ref{trafoforl}). As next step we let~$\hat{V}$ act and then re-transform into the operators for the Peierls subbands in the last step. The calculation shows that only four combinations of currents occur in~$X^{c,s}$ and $Y^{c,s}$, namely, \begin{mathletters} \begin{eqnarray} X^{c,s}(k,q)&=& - f_{+,+}^*(k,q) J_1^{c,s}(q) + f_{+,-}^*(k,q)J_2^{c,s}(q) \\[6pt] Y^{c,s}(k,q)&=& f_{+,-}^*(k,q) J_1^{c,s}(q)+ f_{+,+}^*(k,q) J_2^{c,s}(q) \end{eqnarray} \end{mathletters}% with \begin{mathletters} \label{C5} \begin{eqnarray} J_1^{c,s}(q) &=& \frac{1}{L} \sum_{|p| \leq \pi/(2a)} \left( {\cos(pa) \atop \sin(pa)}\right) \left[ f_{+,+}(j_{-,-}-j_{+,+})+f_{+,-}(j_{+,-}+j_{-,+})\right] \\[6pt] J_2^{c,s}(q) &=& \frac{1}{L} \sum_{|p| \leq \pi/(2a)} \left( {\cos(pa) \atop \sin(pa)}\right) \left[ f_{+,+}(j_{+,-}+j_{-,+})-f_{+,-}(j_{-,-}-j_{+,+})\right] \; . \end{eqnarray} \end{mathletters}% Since the interaction is restricted to nearest-neighbors the global, i.~e., only $q$-dependent currents~$J_{1,2}^{c,s}(q)$ appear in the problem. Equation~(\ref{C2}) becomes \begin{equation} \langle \hat{\jmath}_{\omega >0}\rangle(V) - \langle \hat{\jmath}_{\omega >0}\rangle(V=0) = - \frac{2V{\cal N}_{\perp}}{a} \sum_{|q| \leq \pi/(2a) } \left[ J_1^{c,s}(q)G_1^{c,s}(q)+ J_2^{c,s}(q)G_2^{c,s}(q)\right] \label{C4} \end{equation} with~$G_{1,2}^{c,s}(q)$ given by \begin{eqnarray} G_{1,2}^{c,s} (q) &=& \frac{1}{L} \sum_{|k|\leq \pi/(2a)} \left( {\cos(ka) \atop \sin(ka)}\right) \biggl[ \left( {-f_{+,+}^* \atop f_{+,-}^*}\right)\lambda_{+,+} \left(\frac{1}{\omega-E_{-,-}}+\frac{1}{\omega-E_{+,+}}\right) \nonumber \\[6pt] && \phantom{ \frac{1}{L} \sum_{|k|\leq \pi/(2a)} \left( {\cos(ka) \atop \sin(ka)}\right) \biggl[ } +\left( {f_{+,-}^* \atop f_{+,+}^*} \right) \lambda_{+,-} \left(\frac{1}{\omega-E_{-,+}}+\frac{1}{\omega-E_{+,-}}\right) \biggr] \; . \label{capitalG} \end{eqnarray} The quantities~$G_{1,2}^{c,s} (q)$ are still operator valued objects since they contain $\lambda_{\tau,\tau'}$. Nevertheless these quantities are known. We insert eq.~(\ref{thelambdas}) and use the fact that $|f_{+,+}(-k,q)|^2=|f_{+,+}(k,q)|^2$, $|f_{+,-}(-k,q)|^2=|f_{+,-}(k,q)|^2$, and $f_{+,+}^*(-k,q)f_{+,-}^{\phantom{*}}(-k,q)= f_{+,+}^*(k,q)f_{+,-}^{\phantom{*}}(k,q)$ to show that $G_1^s(q)=G_2^c(q)=0$. We set $G_1^c(q)\equiv G_1(q)$, $G_2^s(q)\equiv G_2(q)$, $J_1^c(q)\equiv J_1(q)$, $J_2^s(q)\equiv J_2(q)$. They can be expressed in terms of the functions~$F_{1,2,3}(q)$ of eq.~(\ref{capitalF}) as \begin{mathletters} \label{CapitalGred} \begin{eqnarray} G_1(q)&=& itea \left[ \hat{x}_q^+ F_1(q) - \hat{x}_{q+\pi/a}^+ F_3(q)\right]\\[6pt] G_2(q)&=& itea \left[ -\hat{x}_q^+ F_3(q) + \hat{x}_{q+\pi/a}^+ F_2(q)\right]\; . \end{eqnarray} \end{mathletters}% It remains to determine~$J_{1,2}(q)$. They can be obtained from their definitions in eq.~(\ref{C5}) and the result from the equations of motion, eq.~(\ref{C1}), \begin{mathletters} \label{C6} \begin{eqnarray} J_{1}(q)+\frac{{\cal A}(\omega)}{c}G_1^{+}(q) &=& V \left( - J_1(q)F_1(q) + J_2(q) F_3 (q)\right) \\[6pt] J_{2}(q)+ \frac{{\cal A}(\omega)}{c}G_2^{+}(q) &=& V\left( - J_2(q) F_2(q) + J_1(q)F_3(q) \right)\; . \end{eqnarray} \end{mathletters}% It is not difficult to invert these equations to obtain the currents explicitly. The result for the real part of the optical conductivity becomes {\arraycolsep=0pt \begin{eqnarray} {\rm Re}\{\overline{\sigma(\omega >0, V, \delta,\eta)} \} &=& {\rm Re}\{\overline{\sigma(\omega >0,\delta,\eta)} \} + \frac{2V{\cal N}_{\perp}}{a\omega} \label{mostimportant} \\[6pt] &&{\rm Im}\Biggl\{ \sum_{|q| \leq \pi/(2a) } \frac{1}{(1+VF_1)(1+VF_2)-(VF_3)^2} \biggl[G_1^{\phantom{+}}G_1^+ + G_2^{\phantom{+}}G_2^+ \nonumber \\[6pt] && \phantom{ {\rm Im}\biggl\{ \sum_{|q| \leq \pi/(2a) } } +V \left( G_1^{\phantom{+}}G_1^+F_2 + G_2^{\phantom{+}}G_2^+F_1 + (G_1^{\phantom{+}}G_2^+ + G_2^{\phantom{+}}G_1^+)F_3 \right)\biggr]\Biggr\} \; . \nonumber \end{eqnarray}} As a last step we have to factorize the products over the functions~$G_{1,2}(q)$. For the Harris-Lange model we use eq.~(\ref{averagexq}). With the help of eq.~(\ref{CapitalGred}) it is not difficult to derive the final result for the average optical conductivity in the Harris-Lange model, eq.~(\ref{thefinalresultHL}). \end{appendix} \newpage \begin{center} {\bf REFERENCES} \end{center} \begin{itemize} \item M.~Abramovitz and I.~A.~Stegun, {\sl Handbook of Mathematical Functions}, (9th~printing, Dover Publications, New York, (1970)). \item L.~Alc\'{a}cer, and A.~Brau and J.-P.~Farges, in: {\sl Organic Conductors}, ed. by J.-P.~Farges (Marcel Dekker, New York, (1994)). \item A.~A.~Aligia and L.~Arrachea, Phys.~Rev.~Lett.~{\bf 73}, 2240 (1994). \item N.~Andrei, {\sl Integrable models in condensed matter physics}, in {\sl Proceedings of the summer school on field theoretical methods (24~August -- 4~September 1992)}, ed.\ by Yu~Lu, (World Scientific, Singapore, (1995)). \item D.~Baeriswyl, P.~Horsch, and K.~Maki, Phys.~Rev.~Lett.~{\bf 60} (C), 70 (1988). \item P.-A.~Bares and F.~Gebhard, Europhys.~Lett.~{\bf 29}, 573, (1995). \item P.-A.~Bares and F.~Gebhard, Journ.\ Low Temp.~Phys.~{\bf 99}, 565 (1995). \item P.-A.~Bares and F.~Gebhard, J.~Phys.~Cond.~Matt.~{\bf 7}, 2285 (1995). \item G.~Beni, P.~Pincus, and T.~Holstein, Phys.~Rev.~B~{\bf 8}, 312 (1973). \item J.~Bernasconi, M.~J.~Rice, W.~R.~Schneider, and S.~Str\"{a}\ss ler, Phys.~Rev.~B~{\bf 12}, 1090 (1975). \item J.~D.~Bjorken and S.~D.~Drell, {\sl Relativistic Quantum Mechanics}, (McGraw-Hill, New York, (1964)). \item J.~de Boer, V.~E.~Korepin, and A.~Schad\-schneider, Phys.~Rev.~Lett.~{\bf 74}, 789 (1995). \item D.~K.~Campbell, J.~T.~Gammel, and E.~Y.~Loh, Phys.~Rev.~B~{\bf 38}, 12043 (1988). \item D.~K.~Campbell, J.~T.~Gammel, and E.~Y.~Loh, Int.~J.~Mod.~Phys.~B~{\bf 3}, 2131 (1989). \item D.~K.~Campbell, J.~T.~Gammel, and E.~Y.~Loh, in: {\sl Interacting Electrons in Reduced Dimensions}, ed.~by D.~Baeriswyl and D.~K.~Campbell (NATO ASI Series~B~{\bf 213}, Plenum Press, New York, (1989)), p.~171. \item D.~K.~Campbell, J.~T.~Gammel, and E.~Y.~Loh, Phys.~Rev.~B~{\bf 42}, 475 (1990). \item P.~G.~J.~van Dongen, Phys.~Rev.~B~{\bf 49}, 7904 (1994). \item P.~G.~J.~van Dongen, Phys.~Rev.~B~{\bf 50}, 14016 (1994). \item F.~H.~L.~E\ss ler and V.~E.~Korepin (ed.), {\sl Exactly Solvable Models of Strongly Correlated Electrons}, (World Scientific, Singapore, (1994)). \item J.-P. Farges (ed.), {\sl Organic Conductors}, (Marcel Dekker, New York, (1994)). \item J.~L.~Fave, in: {\sl Electronic Properties of Polymers}, ed.~by H.~Kuzmany, M.~Mehring, and S.~Roth, (Springer Series in Solid State Sciences~{\bf 107}, Springer, Berlin (1992)). \item H.~Frahm and V.~E.~Korepin, Phys.~Rev.~B~{\bf 42}, 10553 (1990). \item H.~Frahm and V.~E.~Korepin, Phys.~Rev.~B~{\bf 43}, 5653 (1991). \item A.~Fritsch and L.~Ducasse, J.~Physique~I~{\bf 1}, 855 (1991). \item R.~M.~Fye, M.~J.~Martins, D.~J.~Scalapino, J.~Wagner, and W.~Hanke, Phys.~Rev.~B~{\bf 45}, 7311 (1992). \item J.-P.~Galinar, J.~Phys.~C~{\bf 12}, L335 (1979). \item J.~T.~Gammel and D.~K.~Campbell, Phys.~Rev.~Lett.~{\bf 60} (C), 71 (1988). \item F.~Gebhard, K. Bott, M.~Scheidler, P.~Thomas, and S.~W.~Koch, previous article, referred to as~I. \item F.~Gebhard, K. Bott, M.~Scheidler, P.~Thomas, and S.~W.~Koch, forthcoming article, referred to as~III. \item F.~Gebhard, A.~Girndt, and A.~E.~Ruckenstein, Phys.~Rev.~B~{\bf 49}, 10926 (1994) \item F.~Gebhard and A.~E.~Ruckenstein, Phys.~Rev.~Lett.~{\bf 68}, 244 (1992). \item I.~S.~Gradshteyhn and I.~M.~Ryzhik, {\sl A Table of Integrals, Series, and Products}, (Academic Press, New York, (1980)). \item D.~Guo, S.~Mazumdar, S.~N.~Dixit, F.~Kajzar, F.~Jarka, Y.~Kawabe, and N.~Peyghambarian, Phys.~Rev.~B.~{\bf 48}, 1433 (1993). \item F.~Guo, D.~Guo, and S.~Mazumdar, Phys.~Rev.~B.~{\bf 49}, 10102 (1994). \item F.~D.~M.~Haldane, Phys.~Rev.~Lett.~{\bf 67}, 937 (1991). \item E.~R.~Hansen, {\sl A Table of Series and Products}, (Prentice-Hall, Englewood Cliffs, (1975)). \item A.~B.~Harris and R.~V.~Lange, Phys.~Rev.~{\bf 157}, 295 (1967). \item H.~Haug and S.~W.~Koch, {\sl Quantum Theory of the Optical and Electronic Properties of Semiconductors}, (World Scientific, Singapore, (1990)). \item J.~Hubbard, Proc.~R.~Soc. London, Ser.~A~{\bf 276}, 238 (1963). \item N.~Kawakami and S.-K.~Yang, Phys.~Lett.~{\bf A~148}, 359 (1990). \item N.~Kawakami and S.-K.~Yang, Phys.\ Rev.\ Lett.~{\bf 65}, 3063 (1990). \item N.~Kawakami and S.-K.~Yang, Phys.~Rev.~B~{\bf 44}, 7844 (1991). \item S.~Kivelson, W.-P.~Su, J.~R.~Schrieffer, and A.~J.~Heeger, Phys.~Rev.~Lett.~{\bf 58}, 1899 (1987). \item S.~Kivelson, W.-P.~Su, J.~R.~Schrieffer, and A.~J.~Heeger, Phys.~Rev.~Lett.~{\bf 60} (C), 72 (1988). \item D.~J.~Klein, Phys.~Rev.~B~{\bf 8}, 3452 (1973). \item W.~Kohn, Phys.~Rev.~{\bf 133}, A171 (1964). \item E.~H.~Lieb and F.~Y.~Wu, Phys.~Rev.~Lett.~{\bf 20}, 1445 (1968). \item S.~K.~Lyo and J.-P.~Galinar, J.~Phys.~C~{\bf 10}, 1693 (1977). \item S.~K.~Lyo, Phys.~Rev.~B~{\bf 18}, 1854 (1978). \item G.~D.~Mahan, {\sl Many-Particle Physics}, (2nd~edition, Plenum Press, New York (1990)). \item P.~F.~Maldague, Phys.~Rev.~B~{\bf 16}, 2437 (1977). \item S.~Mazumdar and S.~N.~Dixit, Phys.~Rev.~B~{\bf 34}, 3683 (1986). \item S.~Mazumdar and Z.~G.~Soos, Phys.~Rev.~B~{\bf 23}, 2810 (1981). \item F.~Mila, Phys.~Rev.~B~{\bf 52}, 4788 (1995). \item M.~Ogata and H.~Shiba, Phys.~Rev.~B~{\bf 41}, 2326 (1990). \item A.~A.~Ovchinnicov, Zh.~Eksp.~Teor.~Fiz.~{\bf 57}, 2137 (1969) (Sov.~Phys.\ JETP~{\bf 30}, 1160 (1970)). \item A.~Painelli and A.~Girlando, Synth.~Met.~{\bf 27}, A15 (1988). \item A.~Painelli and A.~Girlando, in: {\sl Interacting Electrons in Reduced Dimensions}, ed.~by D.~Baeriswyl and D.~K.~Campbell, (NATO ASI Series~B~{\bf 213}, Plenum Press, New York, (1989)), p.~165. \item A.~Painelli and A.~Girlando, Phys.~Rev.~B~{\bf 39}, 2830 (1989). \item A.~Parola and S.~Sorella, Phys.~Rev.~Lett.~{\bf 60}, 1831 (1990). \item L.~Salem, {\sl Molecular Orbital Theory of Conjugated Systems}, (Benjamin, London, (1966)). \item A.~Schad\-schneider, Phys.~Rev.~B~{\bf 51}, 10386 (1995). \item H.~J.~Schulz, Phys.~Rev.~Lett.~{\bf 64}, 2831 (1990). \item H.~J.~Schulz, in: {\sl Strongly Correlated Electron Systems~II}, (Proceedings of the Adriatico Research Conference and Miniworkshop, Trieste, Italy, 18~June -- 27~July~1990), ed.~by G.~Baskaran, A.~E.~Ruckenstein, E.~Tosatti, and Y.~Lu, (World Scientific, Singapore, (1991)). \item B.~S.~Shastry, S.~S.~Jha, and V.~Singh (ed.), {\sl Exactly Solvable Problems in Condensed Matter and Relativistic Field Theory}, (Lecture Notes in Physics~{\bf 242}, Springer, Berlin, (1985)). \item B.~S.~Shastry and B.~Sutherland, Phys.~Rev.~Lett.~{\bf 65}, 243 (1990). \item W.~T.~Simpson, J.~Am.~Chem.~Soc.~{\bf 73}, 5363 (1951). \item W.~T.~Simpson, J.~Am.~Chem.~Soc.~{\bf 77}, 6164 (1955). \item Z.~G.~Soos and S.~Ramesesha, Phys.~Rev.~B~{\bf 29}, 5410 (1984). \item C.~A.~Stafford and A.~J.~Millis, Phys.~Rev.~B~{\bf 48}, 1409 (1993). \item C.~A.~Stafford, A.~J.~Millis, and B.~S.~Shastry, Phys.~Rev.~B~{\bf 43}, 13660 (1991). \item P.~Tavan and K.~Schulten, J.~Chem.~Phys.~{\bf 85}, 6602 (1986). \item C.-Q.~Wu, X.~Sun, and K.~Nasu, Phys.~Rev.~Lett.~{\bf 59}, 831 (1987). \end{itemize} \begin{figure}[th] \caption{Band structure interpretation of the exact eigenenergies of the Harris-Lange model for U=2W.} \label{HarrisLangedis} \end{figure} \typeout{figure captions} \begin{figure}[th] \caption{Band structure interpretation of the exact eigenenergies of the dimerized Harris-Lange model for U=2W, $\delta=0.2$.} \label{HueckelHarrisLangedis} \end{figure} \begin{figure}[th] \caption{Reduced average optical conductivity, $\overline{\sigma_{\rm red}(\omega >0)}$, in the Harris-Lange model for $U=2W$. A broadening of~$\gamma=0.01W$ has been included.} \label{hl00} \end{figure} \begin{figure}[th] \caption{Reduced average optical conductivity, $\overline{\sigma_{\rm red}(\omega >0,\delta,\eta)}$, in the dimerized Harris-Lange model for $U=2W$, $\delta=0.2$ ($\delta=0.6$), and $\eta=-0.06$. A broadening of~$\gamma=0.01W$ has been included.} \label{hl10} \end{figure} \begin{figure}[th] \caption{Reduced average optical conductivity, $\overline{\sigma_{\rm red}(\omega >0,V)}$, in the extended Harris-Lange model for $U=2W$ and $V=0, W/2, W$. A broadening of~$\gamma=0.01W$ has been included.} \label{hl01} \end{figure} \begin{figure}[th] \caption{Reduced average optical conductivity, $\overline{\sigma_{\rm red}(\omega >0,V,\delta,\eta)}$, in the extended dimerized Harris-Lange model for $U=2W$, $\delta=0.2$, $\eta=-0.06$, $V=0,W/2,W$. A broadening of~$\gamma=0.01W$ has been included.} \label{hl11} \end{figure} \end{document}
proofpile-arXiv_065-472
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{INTRODUCTION} The analysis of primordial nucleosynthesis provides valuable limits on cosmological and particle physics parameters through a comparison between the predicted and inferred primordial abundances of D, \he3, \he4, and \li7. For standard homogeneous big bang nucleosynthesis (HBBN) the predicted primordial abundances of these light-elements are in accord with the value inferred from observation provided that baryon-to-photon ratio ($\equiv \eta$) is between about $2.5 \times 10^{-10}$ and $6 \times 10^{-10}$. This corresponds to an allowed range for the baryon fraction of the universal closure density $\Omega_b^{\rm HBBN}$ (\cite{walker91}; \cite{smith93}; \cite{copi95}; \cite{ScMa95}), \begin{equation} 0.04 \mathrel{\mathpalette\fun <} \hbox{$\Omega_b$}^{\rm HBBN}\,h_{50}^{2} \mathrel{\mathpalette\fun <} 0.08 , \label{eq:1} \end{equation} where $\eta = 6.6 \times 10^{-9} \Omega_b\,h_{50}^2$. The lower limit on $\hbox{$\Omega_b$}^{\rm HBBN}$ arises mainly from the upper limit on the deuterium plus $^3$He abundance (\cite{yang84}; \cite{walker91}; \cite{smith93}), and the upper limit to $\hbox{$\Omega_b$}$ arises from the upper limit on the \he4 mass fraction $\hbox{$Y_{\rm p}$}$ and/or the deuterium abundance D/H $\ge 1.2 \times 10^{-5}$ (\cite{linsky93},\ \cite{linsky95}). Here, $h_{50}$ is the Hubble constant in units of 50 km s $^{-1}$ Mpc$^{-1}$. The fact that this range for $\hbox{$\Omega_b$}\,h_{50}^2$ is so much greater than the current upper limit to the contribution from luminous matter $\hbox{$\Omega_b$}^{Lum} \mathrel{\mathpalette\fun <} 0.01$ (see however \cite{jedamzik95}) is one of the strongest arguments for the existence of baryonic dark matter. Over the years HBBN has provided strong support for the standard, hot big bang cosmological model as mentioned above. However, as the astronomical data have become more precise in recent years, a possible conflict between the predicted abundances of the light element isotopes from HBBN and the abundances inferred from observations has been suggested (\cite{olivestei95}; \cite{steigman96a}; \cite{turner96}; \cite{hata96}; see also \cite{hata95}). There is now a good collection of abundance information on the \he4 mass fraction, \hbox{$Y_{\rm p}$}, O/H, and N/H in over 50 extragalactic HII regions (\cite{pagel92}; \cite{pagel93}; \cite{izatov94}; \cite{skillman95}). In an extensive study based upon these observations, the upper limit to $\eta$ from the observed \he4 abundance was found to be $\sim 3.5 \times 10^{-10}$ (\cite{olivestei95}; \cite{olive96}) when a systematic error in $\hbox{$Y_{\rm p}$}$ of $\Delta Y_{sys} = 0.005$ is adopted. Recently, it has been recognized that the $\Delta Y_{sys}$ may even be factor of 2 or 3 larger (\cite{thuan96}; \cite{copi95}; \cite{ScMa95}; \cite{sasselov95}), making the upper limit to $\eta$ as large as $7 \times 10^{-10}$. On the other hand, the lower bound to $\eta$ has been derived directly from the upper bound to the combined abundances of D and \he3. This is because it is believed that deuterium is largely converted into \he3 in stars; the lower bound then applies if, as has generally been assumed, a significant fraction of \he3 survives stellar processing (\cite{walker91}). However, there is mounting evidence that low mass stars destroy \he3 (\cite{wasserburg95}; \cite{charbonnel95}), although it is possible that massive stars produce \he3. Therefore, the uncertainties of chemical evolution models render it difficult to infer the primordial deuterium and \he3 abundances by using observations of the present interstellar medium (ISM) or from the solar meteoritic abundances. Recent data and analysis lead to a lower bound of $\eta\,\mathrel{\mathpalette\fun >}\,3.5\,\times\,10^{-10}$ on the basis of D and \he3 (\cite{dearborn96}; \cite{hata96}; \cite{steigman96a}; \cite{steigman95}), if the fraction of \he3 that survives stellar processing in the course of galactic evolution exceeds $1/4$. This poses a potential conflict between the observation ($\hbox{$Y_{\rm p}$}$ with low $\Delta Y_{sys}$, D) and HBBN. In this context, possible detections (\cite{songaila94}; \cite{carswell94}; \cite{carswell96}; \cite{tytler94}; \cite{tytler96}; \cite{rugers96a},1996b; \cite{wampler96}) of an isotope-shifted Lyman-$\alpha$ absorption line at high redshift ($z\, \mathrel{\mathpalette\fun >}\,3$) along the line of sight to quasars are of considerable interest. Quasar absorption systems can sample low metallicity gas at early epochs where little destruction of D should have occurred. Thus, they should give definitive measurements of the primordial cosmological D abundance. A very recent high resolution detection by Rugers \& Hogan (1996a) suggests a ratio D/H of \begin{equation} \rm{D}/\rm{H} = 1.9 \pm 0.4 \times 10^{-4}. \label{eq:2} \end{equation} This result is consistent with the estimates made by Songaila et al.~(1994) and Carswell et al.~(1994), using lower resolution. It is also similar to that found recently in another absorption system by Wampler et al.~(1996), but it is inconsistent with high resolution studies in other systems at high redshift (Tytler, Fan \& Burles 1996; Burles \& Tytler 1996) and with the local observations of D and \he3 in the context of conventional models of stellar and Galactic evolution (\cite{edmunds95}; \cite{gloeckler96}). If the high value of D/H is taken to be the primordial abundance, then the consistency between the observation and HBBN is recovered and the allowed range of $\hbox{$\Omega_b$}$ inferred from HBBN changes to $\hbox{$\Omega_b$}^{\rm HBBN}\,h_{50}^{2} = 0.024 \pm 0.002$ (\cite{jedamzik94a}; \cite{krauss94a}; \cite{vangioni95}). In this case, particularly if $h_{50}$ is greater than $\sim 1.5$, the big bang prediction could be so close to the baryonic density in luminous matter that little or no baryonic dark matter is required (\cite{persic92}; \cite{jedamzik95}). This could be in contradiction with observation, particularly if the recently detected microlensing events (Alcock et al.~1993, 1994, 1995abc; \cite{aubourg93}) are shown to be baryonic. This low baryonic density limit would also be contrary to evidence (\cite{white93}; \cite{white95}) that baryons in the form of hot X-ray gas may contribute a significant fraction of the closure density The observations by Tytler, et al.~(1996) and Burles \& Tytler (1996) yield a low value of D/H. Their average abundance is \begin{equation} \rm{D}/\rm{H} = 2.4 \pm 0.9 \times 10^{-5}, \label{eq:3} \end{equation} with $\pm 2\sigma$ statistical error and $\pm 1\sigma$ systematic error. This value is consistent with the expectations of local galactic chemical evolution. However, this value would imply an HBBN helium abundance of $\hbox{$Y_{\rm p}$} = 0.249 \pm 0.003$ which is only marginally consistent with the observationally inferred $\hbox{$Y_{\rm p}$}$ even if the high $\Delta Y_{sys}$ is adopted. With this in mind, it is worthwhile to consider alternative cosmological models. One of the most widely investigated possibilities is that of an inhomogeneous density distribution at the time of nucleosynthesis. Such studies were initially motivated by speculation (\cite{witten84}; \cite{applegate85}) that a first order quark-hadron phase transition (at $T \sim 100 \mbox{~MeV}$) could produce baryon inhomogeneities as baryon number was trapped within bubbles of shrinking quark-gluon plasma. In previous calculations using the baryon inhomogeneous big bang nucleosynthesis (IBBN) model, it has been usually assumed that the geometry of baryon density fluctuations is approximated by condensed spheres. Such geometry might be expected to result from a first order QCD phase transition in the limit that the surface tension dominated the evolution of shrinking bubbles of quark-gluon plasma. However, the surface tension may not be large (\cite{kajantie90}, 1991, 1992) during the QCD transition, which could lead to a "shell" geometry or the development of dendritic fingers (Freese \& Adams 1990). Furthermore, such fluctuations might have been produced by a number of other processes operating in the early universe (cf. \cite{malaney93}), for which other geometries may be appropriate, e.g. strings, sheets, etc. Thus, the shapes of any cosmological baryon inhomogeneities must be regarded as uncertain. The purpose of this paper is, therefore, to explore the sensitivity of the predicted elemental abundances in IBBN models to the geometry of the fluctuations. We consider here various structures and profiles for the fluctuations in addition to condensed spheres. Mathews et al.~ (1990, 1994, 1996) found that placing the fluctuations in spherical shells rather than condensed spheres allowed for lower calculated abundances of \he4 and \li7 for the same $\hbox{$\Omega_b$}$, and that a condensed spherical geometry is not necessarily the optimum. Here we show that a cylindrical geometry also allows for an even higher baryonic contribution to the closure density than that allowed by the usually adopted condensed sphere. It appears to be a general result that shell geometries allow for a slightly higher baryon density. This we attribute to the fact that, for optimum parameters, shell geometries involve a larger surface area to volume ratio and hence more efficient neutron diffusion. An important possible consequence of baryon inhomogeneities at the time of nucleosynthesis may be the existence of unique nucleosynthetic signatures. Among the possible observable signatures of baryon inhomogeneities already pointed out in previous works are the high abundances of heavier elements such as beryllium and boron (\cite{boyd89}; \cite{kajino90a}; \cite{malaney89a}; Terasawa \& Sato 1990; \cite{kawano91}), intermediate mass elements (\cite{kajino90b}), or heavy elements (\cite{malaney88}; \cite{applegate88}; \cite{rauscher94}). Such possible signatures are also constrained, however, by the light-element abundances. It was found in several previous calculations that the possible abundances of synthesized heavier nuclei was quite small (e.g., \cite{alcock90}; Terasawa \& Sato 1990; \cite{rauscher94}). We find, however, that substantial production of heavier elements may nevertheless be possible in IBBN models with cylindrical geometry. \section{BARYON DENSITY INHOMOGENEITIES} After the initial suggestion (Witten 1985) of QCD motivated baryon inhomogeneities it was quickly realized (\cite{applegate85}; \cite{applegate87}) that the abundances of primordial nucleosynthesis could be affected. A number of papers have addressed this point (\cite{alcock87}; \cite{applegate87}, 1988; \cite{fuller88}; \cite{kurki88}, 1990; \cite{terasawa89a}, \cite{terasawa89b}, \cite{terasawa89c}, \cite{terasawa90}; \cite{kurki89}, \cite{kurki90a}; \cite{mathews90}, \cite{mathews93b}; \cite{mathews96}; \cite{jedamzik94a}; \cite{jedamzik95}; \cite{thomas94}; \cite{rauscher94}). Most recent studies in which the coupling between the baryon diffusion and nucleosynthesis has been properly accounted for (e.g., \cite{terasawa89a}, \cite{terasawa89b}, \cite{terasawa89c}, \cite{terasawa90}; \cite{kurki89}, \cite{kurki90a}; \cite{mathews90}, \cite{mathews93b}; \cite{jedamzik94a}; \cite{thomas94}) have concluded that the upper limit on $\hbox{$\Omega_b$}\,h^2$ is virtually unchanged when compared to the upper limit on $\hbox{$\Omega_b$}\,h^2$ derived from standard HBBN. It is also generally believed (e.g. \cite{vangioni95}) that the same holds true if the new high D/H abundance is adopted. However, in the previous studies, it was usually assumed that a fluctuation geometry of centrally condensed spheres produces the maximal impact on nucleosynthesis. Here we emphasize that condensed spheres are not necessarily the optimal nor the most physically motivated fluctuation geometry. Several recent lattice QCD calculations (\cite{kajantie90}, 1991, 1992; \cite{brower92}) indicate that the surface tension of nucleated hadron bubbles is relatively low. In this case, after the hadron bubbles have percolated, the structure of the regions remaining in the quark phase may not form spherical droplets but rather sheets or filaments. We do note that the significant effects on nucleosynthesis may require a relatively strong first order phase transition and sufficient surface tension to generate an optimum separation distance between baryon fluctuations (\cite{fuller88}). However, even if the surface tension is low, the dynamics of the coalescence of hadron droplets may lead to a large separation between regions of shrinking quark-gluon plasma. Furthermore, even though lattice QCD has not provided convincing evidence for a strongly first order QCD phase transition (e.g., \cite{fukugita91}), the order of the transition must still be considered as uncertain (\cite{gottlieb91}; \cite{petersson93}). It depends sensitively upon the number of light quark flavors. The transition is first order for three or more light flavors and second order for two. Because the $s$ quark mass is so close to the transition temperature, it has been difficult to determine the order of transition. At least two recent calculations (\cite{iwasaki95}; \cite{kanaya96}) indicate a clear signature of a first order transition when realistic $u, d, s$ quark masses are included, but others indicate either second order or no phase transition at all. In addition to the QCD phase transition, there remain a number of alternative mechanisms for generating baryon inhomogeneities prior to the nucleosynthesis epoch (cf.~\cite{malaney93}), such as electroweak baryogenesis (\cite{fuller94}), inflation-generated isocurvature fluctuations (\cite{dolgov93}), and kaon condensation (\cite{nelson90}). Cosmic strings might also induce baryon inhomogeneities through electromagnetic (\cite{malaney89b}) or gravitational interactions. Since the structures, shapes, and origin of any baryon inhomogeneities are uncertain, a condensed spherical geometry is not necessarily the most physically motivated choice. Indeed, we will show that a condensed spherical geometry is also not necessarily the optimum to allow for the highest values for $\hbox{$\Omega_b$}$ while still satisfying the light-element abundance constraints. Here we consider the previously unexplored cylindrical geometry. String geometries may naturally result from various baryogenesis scenarios such as superconducting axion strings or cosmic strings. Also, the fact that QCD is a string theory may predispose QCD-generated fluctuations to string-like geometry (\cite{kajino93}; \cite{tassie93}). Hence, cylindrical fluctuations may be a natural choice. \section{OBSERVATIONAL CONSTRAINTS} We adopt the following constraints on the observed helium mass fraction $\hbox{$Y_{\rm p}$}$ and \li7 taken from Balbes et al.~(1993), Schramm \& Mathews (1995), Copi et al.~(1995) and Olive (1996): \begin{equation} 0.226 \leq \hbox{$Y_{\rm p}$} \leq 0.247, \label{eq:4} \end{equation} \begin{equation} 0.7 \times 10^{-10} \leq \li7/\rm{H} \leq 3.5 \times 10^{-10}. \label{eq:5} \end{equation} This primordial $\he4$ abundance constraint includes a statistical uncertainty of $\pm 0.003$ and possible systematic errors as much as $+0.01/-0.005$ with central value of $0.234$. A recent reinvestigation (with new data) of the linear regression method for estimating the primordial $\he4$ abundance has called into question the systematic uncertainties assigned to $\hbox{$Y_{\rm p}$}$ (\cite{izatov96}). Our adopted upper limit to $\hbox{$Y_{\rm p}$}$ of Eq.~(\ref{eq:4}) is essentially equal to the limit derived in their study with $1\sigma$ statistical error. The upper limit to the lithium abundance adopted here includes the systematic increase from the model atmospheres of Thorburn (1994) and the possibility of as much as a factor of 2 increase due to stellar destruction. This is consistent with the recent observations of \li6 in halo stars (\cite{smith92}; \cite{hobbs94}). We note that recent discussion of model atmospheres (\cite{kurucz95}) suggests that as much as an order magnitude upward shift in the primordial lithium abundance could be warranted due to the tendency of one-dimensional models to underestimate the ionization of lithium. Furthermore, a recent determination of the lithium abundance in the globular cluster M92 having the metallicity [Fe/H] = -2.25 has indicated that at least one star out of seven shows [Li] =\,12\,+\,log(Li/H) $\approx 2.5$ (\cite{boesgaard96b}). Since the abundance measurement of the globular cluster stars is more reliable than that of field stars, this detection along with the possible depletion of lithium in stellar atmospheres suggests that a lower limit to the primordial abundance is $ 3.2 \times 10^{-10} \leq \li7/\rm{H}$. There also remains the question as to why several stars which are in all respects similar to the other stars in the Population~II `lithium plateau', are so lithium rich or lithium deficient (\cite{deliyannis96}; \cite{boesgaard96a}, 1996b). Until this is clarified, it may be premature to assert that the Population~II abundance of lithium reflects the primordial value. The primordial abundance may instead correspond to the much higher value observed in Population~I stars which has been depleted down to the Population~II lithium plateau. The observational evidence (Deliyannis, Pinsonneault \& Duncan 1993) for a $\pm\,25\,\%$ dispersion in the Population~II lithium plateau is consistent with this hypothesis (\cite{deliyannis93}; \cite{charbonnel95}; \cite{steigmanli7}). Rotational depletion was studied in detail by Pinsonneault et al.~ (1992) who note that the depletion factor could have been as large as $10$. Chaboyer and Demarque (1994) also demonstrated that models incorporating both rotation and diffusion provide a good match to the observed \li7 depletion with decreasing temperature in Population~II stars and their model indicated that the initial lithium abundance could have been as high as $\li7/\rm{H} = 1.23 \pm\,0.28 \times 10^{-9}$. A recent study (\cite{ryan96}), which includes new data on 7 halo dwarfs, fails to find evidence of significant depletion through diffusion, although other mechanisms are not excluded. For example, stellar wind-driven mass loss could deplete a high primordial lithium abundance of down to the Population~II value [Eq.~(\ref{eq:5})] in a manner consistent with \li6 observations (\cite{vauclair95}). Furthermore, it could be possible (\cite{yoshii95}) that some of the \li6 is the result of more recent accretion of interstellar material that could occur as halo stars episodically plunge through the disk. Such a process could mask the earlier destruction of lithium. For comparison, therefore, we adopt a conservative upper limit on the primordial lithium abundance of \begin{equation} \li7/\rm{H} < 1.5 \times 10^{-9}. \label{eq:6} \end{equation} Finally, the primordial abundance of deuterium is even harder to clarify since it is easily destroyed in stars (at temperatures exceeding about $6 \times 10^{5}$K). Previously, limits on the deuterium (and\ also the\ \he3) abundances have been inferred from their presence in presolar material (e.g.,~\cite{walker91}). It is also inferred from the detection in the local interstellar medium (ISM) through its ultraviolet absorption lines in stellar spectra (\cite{mccullough92}; \cite{linsky93},\ 1995). The limit from ISM data is consistent with that from abundances in presolar material. It has been argued that there are no important astrophysical sources of deuterium (\cite{epstein76}) and ongoing observational attempts to detect signs of deuterium synthesis in the Galaxy are so far consistent with this hypothesis (see~\cite{pasachoff89}). If this is indeed so, then the lowest D abundance observed today should provide a lower bound to the primordial abundance. Recent precise measurements by Linsky et al.~ (1995, 1993) using the {\it Hubble Space Telescope} implies \begin{equation} \rm{D}/\rm{H} > 1.2 \times 10^{-5}. \label{eq:7} \end{equation} We adopt this as a lower limit to the primordial deuterium abundance for the purposes of exploring the maximal cosmological impact from IBBN. In addition, we consider the two possible detections of the deuterium abundance along the line of sight to high red shifted quasars, Eqs.~(\ref{eq:2}) and (\ref{eq:3}) as possible limits. In order to derive a lower limit to $\hbox{$\Omega_b$}\,h_{50}^2$, it is useful to consider the sum of deuterium plus \he3. In the context of a closed-box instantaneous recycling approximation, it is straightforward (\cite{olive90}) to show that the sum of primordial deuterium and \he3 can be written \begin{equation} y_{23p} \le A_\odot^{(g_3 - 1)}y_{23\odot} \biggl({ X_\odot \over X_p} \biggr) \label{eq:8} \end{equation} where $A_\odot$ is the fraction of the initial primordial deuterium still present when the solar system formed, $g_3$ is the fraction of \he3 that survives incorporation into a single generation of stars, $y_{23\odot}$ is the presolar value of [D$+$\he3]/H inferred from the gas rich meteorites, and $X_\odot/X_p$ is the ratio of the presolar hydrogen mass fraction to the primordial value. These factors together imply an upper limit (\cite{walker91}; \cite{copi95}) of \begin{equation} y_{23p} \le 1.1 \times 10^{-4}. \label{eq:9} \end{equation} \section{CALCULATIONS} The calculations described here are based upon the coupled diffusion and nucleosynthesis code of Mathews et al.~(1990), but with a number of nuclear reaction rates updated and the numerical diffusion scheme modified to accommodate cylindrical geometry. We also have implemented an improved numerical scheme which gives a more accurate description of the effects of proton and ion diffusion, and Compton drag at late times. Although our approach is not as sophisticated as that of Jedamzik et al.~(1994a), it produces essentially the same results for the parameters employed here. We have also included all of the new nuclear reaction rates summarized in Smith et al.~(1993) as well as those given in Thomas et al.~(1993). We obtain the same result as Smith et al.~(1993) using these rates and homogeneous conditions in our IBBN model Calculations were performed in a cylindrical geometry both with the high density regions in the center (condensed cylinders), and with the high density regions in the outer zone of computation (cylindrical shells). Similarly, calculations were made in a spherical geometry with the high density regions in the center (condensed spheres) and with the high density region in the outer zones of computation (spherical shells). In the calculations, the fluctuations are resolved into 16 zones of variable width as described by Mathews et al.~(1990). We assumed three neutrino flavors and an initially homogeneous density within the fluctuations. Such fluctuation shapes are the most likely to emerge, for example, after neutrino-induced expansion (\cite{jedamzik94}). We use a neutron mean life-time of $\tau_{n} = 887.0$ (\cite{particle94}). In addition to the cosmological parameter, $\hbox{$\Omega_b$}$ and fluctuation geometry, there remain three parameters to specify the baryon inhomogeneity. They are: $R$, the density contrast between the high and low-density regions; $f_{v}$, the volume fraction of the high-density region; and $r$, the average separation distance between fluctuations. \section{RESULTS} The parameters $R$ and $f_{v}$ were optimized to allow for the highest values for $\hbox{$\Omega_b$}\,h_{50}^{2}$ while still satisfying the light-element abundance constraints. For fluctuations represented by condensed spheres, optimum parameters are $R \sim 10^{6}$ and $f_v^{1/3} \sim 0.5$ (\cite{mathews96}). For other fluctuation geometries, we have found that optimum parameters are: \begin{mathletters} \begin{eqnarray*} R &\sim & 10^{6}; \qquad \qquad \mbox{for all fluctuation geometries} \\ \\ f_{v}^{1/3} &\sim & 0.19; \qquad \quad \mbox{~~} \mbox{for spherical shells} \\ \\ f_{v}^{1/2} &\sim &\left\{ \begin{array}{rl} 0.5;& \quad \quad \mbox{for condensed cylinders} \\ 0.15;& \quad \quad \mbox{for cylindrical shells}, \end{array}\right. \end{eqnarray*} \mbox{although there is not much sensitivity to $R$ once $R\,\mathrel{\mathpalette\fun >}\, 10^3$.} \end{mathletters} Regarding $f_v$, we have written the appropriate length scale of high density regions, i.e. $f_v^{1/3}$ and $f_v^{1/2}$ for the spherical and cylindrical fluctuation geometries, respectively. The variable parameters in the calculation are then the fluctuation cell radius $r$, and the total baryon-to-photon ratio $\eta$ (or $\hbox{$\Omega_b$}\,h_{50}^{2}$). \subsection{Constraints on \hbox{$\Omega_b$}$h_{50}^2$} Figures~\ref{fig:1}, \ref{fig:2}, \ref{fig:3a}, and \ref{fig:4a} show contours of allowed parameters in the $r$ versus $\eta$ and $r$ versus $\hbox{$\Omega_b$}\,h_{50}^{2}$ plane for the adopted light-element abundance constraints of Eqs.~(\ref{eq:4}) - (\ref{eq:6}) and for a possible Lyman-$\alpha$ D/H of Eqs.~(\ref{eq:2}) and (\ref{eq:3}), for the condensed sphere, spherical shell, condensed cylinder, and cylindrical shell fluctuation geometries, respectively. The fluctuation cell radius $r$ is given in units of meters for a comoving length scale fixed at a temperature of $kT = 1 \mbox{~MeV}$. Both of the possible \li7 limits, Eqs.~(\ref{eq:5}) and (\ref{eq:6}) which we have discussed above, are also drawn as indicated. In order to clearly distinguish the two abundance constraints, we use the single and double-cross hatches for the regions allowed by the adopted lower (Eq.~(\ref{eq:5})) and higher (Eq.~(\ref{eq:6})) limits to the \li7 primordial abundance. Even in the IBBN scenario, if the low D/H of Eq.~(\ref{eq:3}) (\cite{burles96}) is adopted as primordial, this range for D/H appears to be compatible with the \li7 abundance only when a higher (Population~I) primordial \li7 abundance limit is adopted, except for a very narrow region of $\eta \sim 6 \times 10^{-10}$ and $r \leq 10^{2}\,m$. This conclusion remains unchanged for any other fluctuation geometries. Therefore, the acceptance of the low (Burles \& Tytler 1996) value of D/H would strongly suggest that significant depletion of \li7 has occurred. In contrast, adoption of the high D/H of Eq.~(\ref{eq:2}) (\cite{rugers96a}) as primordial allows the concordance of all light-elements. The upper limits to $\eta$ and $\hbox{$\Omega_b$}\,h_{50}^{2}$ are largely determined by D and $\li7$. The concordance range for the baryon density is comparable to that for HBBN for small separation distance $r$. However, there exist other regions of the parameter space with optimum separation distance, which roughly corresponds to the neutron diffusion length during nucleosynthesis (\cite{mathews90}), with an increased maximum allowable value of the baryonic contribution to the closure density to $\hbox{$\Omega_b$}\,h_{50}^{2} \leq 0.05$ for the cylindrical geometry, as displayed in Fig.~\ref{fig:4a}. This is similar to the value for spherical shells as shown in Mathews et al.~(1996) and also in Fig.~\ref{fig:2} in the present work. The condensed sphere limits, however, are essentially unchanged from those of the HBBN model. If the primordial \li7 abundance could be as high as the upper limit of $\mbox{Li/H} \leq 1.5 \times 10^{-9}$, the maximum allowable value of the baryonic content for the condensed sphere would increase to $\hbox{$\Omega_b$}\,h_{50}^{2} \leq 0.08$, with similar values for the spherical shell (\cite{mathews96}). For both the condensed cylinders and cylindrical shells, the upper limits could be as high as $\hbox{$\Omega_b$}\,h_{50}^{2} \leq 0.1$ as shown in Figs.~\ref{fig:3a} and \ref{fig:4a}. These higher upper limits relative to those of the HBBN are of interest since they are consistent with the inferred baryonic mass in the form of hot X-ray gas (\cite{white93}; \cite{white95}) in dense galactic clusters. The acceptance of this consistence, as noted above, requires the significant stellar depletion of \li7. In Figures~\ref{fig:3b} and \ref{fig:4b}, we also show contours for the condensed cylinder and cylindrical shell geometries, respectively, but this time with the conventional light-element constraints of Eqs.~(\ref{eq:4}), (5), (7), and (9) as indicated. Since the results for the condensed sphere and spherical shell geometries with this set of the conventional abundance constraints have already been discussed by Mathews et al.~(1996), we do not show those contours here. The cylindrical shell geometry of the present work gives the highest allowed value of $\hbox{$\Omega_b$}\,h_{50}^{2}$. Figure~\ref{fig:4b} shows that the upper limits to $\eta$ and $\hbox{$\Omega_b$}\,h_{50}^{2}$ are largely determined by $\hbox{$Y_{\rm p}$}$ and $\li7$. The upper limits for a cylindrical shell geometry could be as high as $\hbox{$\Omega_b$}\,h_{50}^{2} \leq 0.13$ with similar results for the spherical shell geometry (\cite{mathews96}). A high primordial lithium abundance would increase the allowable baryonic content to as high as $\hbox{$\Omega_b$}\,h_{50}^{2} \leq 0.2$. The reason that shell geometries allow for higher baryon densities we attribute to more efficient neutron diffusion which occurs when the surface area to volume area is increased. This allows for more initial diffusion to produce deuterium, and more efficient back diffusion to avoid over producing \li7. \subsection{Observational Signature} The production of beryllium and boron as well as lithium in IBBN models can be sensitive to neutron diffusion. Therefore, their predicted abundances are sensitive to not only the fluctuation parameter $r$, $R$, and $f_{v}$ but also the fluctuation geometry (\cite{boyd89}; \cite{malaney89a}; \cite{kajino90a}; Terasawa \& Sato 1990). Figures~\ref{fig:5} - \ref{fig:7} show the contours of the calculated abundances for lithium, beryllium and boron, respectively in the $r$ versus $\eta$ (and $r$ versus $\hbox{$\Omega_b$}~h_{50}^{2}$) plane. the shaded region depict is allowed values of $r$ and $\eta$ from the light element abundance constraints [cf. Fig.~\ref{fig:4b}] for a cylindrical shell fluctuation geometry. The contour patterns of lithium (Fig.~\ref{fig:5}) and boron (Fig.~\ref{fig:7}) abundances are very similar, whereas there is no similarity found between lithium (Fig.~\ref{fig:5}) and beryllium (Fig.~\ref{fig:6}) abundances. In order to understand the similarities and differences among these three elemental abundances, we show in Figs.~\ref{fig:8} and \ref{fig:9} the decompositions of the A = 7 abundance into \li7 and \be7 and the boron abundance into $^{10}$B and $^{11}$B. These Figures show also the dependence of the predicted LiBeB abundances in IBBN on the scale of fluctuations for a cylindrical shell geometry with fixed $\hbox{$\Omega_b$}\,h_{50}^{2} = 0.1$. This value of $\hbox{$\Omega_b$}\,h_{50}^{2}$ corresponds to a typical value in the allowable range of $\eta$ in Fig.~\ref{fig:4b}, which optimizes the light element abundance constraints, even satisfying the lower \li7 abundance limit of Eq.~(\ref{eq:5}). The fluctuation parameters $f_{v}$ and $R$ are the same as in Fig.~\ref{fig:4b}. Once the baryonic content $\hbox{$\Omega_b$}$ is fixed, the only variable parameter is the separation distance, $r$. As can be seen in Fig.~\ref{fig:8}, as the separation $r$ increases, neutron diffusion plays an increasingly important role in the production of $t$ and, by the \he4($t, \gamma$)\li7 reaction. It works maximally around $r \sim 10^{4}$~m, which is the typical length scale of neutron diffusion at $kT = 1 \mbox{~MeV}$. A similar behavior is observed in the \li7($t, n$)\be9 reaction. This reaction produces most of the \be9 in neutron rich environments where $t$ and \li7 are abundant, as was first pointed out by Boyd and Kajino (1989). At other separation distances $r$ in a $\hbox{$\Omega_b$}\,h_{50}^{2} = 0.1$ model, most of the A = 7 nuclides are created as \be7 by the \he4(\he3, $\gamma$)\be7 reaction. In the limit of $r$ = horizon scale, the nucleosynthesis products are approximately equal to the sum of those produced in the proton-rich and neutron-rich zones separately (\cite{jedamzik94b}). The predominant contribution from the proton-rich zones makes the \be7 abundance almost constant at larger $r$, while both \li7 and \be9 decrease as $r$ increases toward the horizon at any separation distance. Figure~\ref{fig:9} shows that $^{11}$B is a predominant component of the total boron abundance at any separation distance. This is true for almost all values $\hbox{$\Omega_b$}\,h_{50}^{2}$. It has been pointed out (\cite{malaney88}; \cite{applegate88}; \cite{kajino90a}) that most $^{11}$B is produced by the \li7($n, \gamma$)\li8($\alpha, n$)$^{11}$B reaction sequence in neutron-rich environments at relatively early times when most of the other heavier nuclides are made. Recent measurements of the previously unmeasured \li7($\alpha$,n)$^{11}$B reaction cross section (\cite{boyd92}; \cite{gu95}; \cite{boyd96}) at the energies of cosmological interest have removed the significant ambiguity in the calculated $^{11}$B abundance due to this reaction. The factor of two discrepancy among several different measurements of the reaction cross section for \li7($n, \gamma$)\li8 was also resolved by the new measurement (\cite{nagai91}). The \li7($\alpha, \gamma$)$^{11}$B reaction also makes an appreciable but weaker contribution to the production of $^{11}$B in the neutron-rich environment. In the proton-rich environment, on the other hand, the \be7($\alpha, \gamma$)$^{11}$C reaction contributes largely to the production of $^{11}$C which beta decays to $^{11}$B in 20.39 min. These facts explain why the contour patterns of the lithium and boron abundances in Figs.~\ref{fig:5} and \ref{fig:7} look very similar. It is conventional in the literature to quote the beryllium and boron abundance relative to H $=10^{2}$. Hence, one defines the quantity [X] $= 12+\mbox{log(X/H)}$. In cylindrical shell fluctuation geometry the beryllium abundance can take the value of $\mbox{[Be]} \sim -3$ while still satisfying all of the light-element abundance constraints and the Population~II lithium abundance constraint (Figs.~\ref{fig:5} and \ref{fig:6}). This abundance is higher by three orders magnitude than that produced in the HBBN model with conventional light-element abundance constraints. This result is contrary to a recent result with the condensed sphere geometry and for a more restricted parameter space (\cite{thomas94}). Recent beryllium observation of Population~II stars (\cite{rebolo88}; \cite{ryan90},\ 1992; \cite{ryan96a}; \cite{gilmore92a},\ 1992b; \cite{boesgaard93}; \cite{boesgaard94},\ 1996a,b) have placed the upper limit on the primordial \be9 abundance to $\mbox{[Be]} \sim -2$, one order magnitude greater than the beryllium abundance in the IBBN cylindrical model. The calculated boron abundance at the optimum separation distance is essentially equal to the value of the HBBN model. However, a high primordial lithium abundance would increase the upper limit to $\hbox{$\Omega_b$}\,h_{50}^{2}$. In this case, the boron abundance could be one or two orders magnitude larger than that of the HBBN model (Fig.~\ref{fig:7}). \section{CONCLUSIONS} We have reinvestigated the upper limit to $\eta$ and $\hbox{$\Omega_b$}\,h_{50}^{2}$ in inhomogeneous primordial nucleosynthesis models. We have considered effects of various geometries. In particular, for the first time we consider cylindrical geometry. We have also incorporated recently revised light-element abundance constraints including implications of the possible detection (\cite{songaila94}; \cite{carswell94}; \cite{carswell96}; \cite{tytler94}; \cite{tytler96}; \cite{rugers96a}, 1996b; \cite{wampler96}) of a high deuterium abundance in Lyman-$\alpha$ absorption systems. We have shown that with low primordial deuterium (\cite{tytler94}; \cite{tytler96}), significant depletion of \li7 is required to obtain concordance between predicted light-element abundance of any model of BBN and the observationally inferred primordial abundance. If high primordial deuterium (\cite{rugers96a}) is adopted (Eq.~(\ref{eq:2})), there is a concordance range which is largely determined by D/H, and the upper limit to $\hbox{$\Omega_b$}\,h_{50}^{2}$ is 0.05. However, with the presently adopted (Eqs.~(\ref{eq:4}), (\ref{eq:6}), (\ref{eq:7}), (\ref{eq:9})) light-element abundance constraints (\cite{ScMa95}; \cite{copi95}; \cite{olive96}), values of $\hbox{$\Omega_b$}\,h_{50}^{2}$ as large as 0.2 are possible in IBBN models with cylindrical-shell fluctuation geometry. We have also found that significant beryllium and boron production is possible in IBBN models without violating the light element abundance constraints. The search for the primordial abundance of these elements in low metallicity stars could, therefore, be a definitive indicator of the presence or absence of cylindrical baryon inhomogeneities in the early universe.
proofpile-arXiv_065-473
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} Recently much attention has been given to lower dimensional gauge theories. Such remarkable results as the chiral symmetry breaking \cite{1}, quantum Hall effect \cite{2}, spontaneously broken Lorentz invariance by the dynamical generation of a magnetic field \cite{3}, and the connection between non-perturbative effects in low-energy strong interactions and QCD$% _{2}$ \cite{4}, show the broad range of applicability of these theories. In particular, 2+1 dimensional gauge theories with fractional statistics -anyon systems \cite{4a}- have been extensively studied. One main reason for such an interest has been the belief that a strongly correlated electron system in two dimensions can be described by an effective field theory of anyons \cite{5}, \cite{5a}. Besides, it has been claimed that anyons could play a basic role in high-T$_{C}$ superconductivity \cite{5a}-\cite{6b}. It is known \cite{a} that a charged anyon system in two spatial dimensions can be modeled by means of a 2+1 dimensional Maxwell-Chern-Simons (MCS) theory. An important feature of this theory is that it violates parity and time-reversal invariance. However, at present no experimental evidences of P and T violation in high-T$_{C}$ superconductivity have been found. It should be pointed out, nevertheless, that it is possible to construct more sophisticated P and T invariant anyonic models\cite{6a}. In any case, whether linked to high-T$_{C}$ superconductivity or not, the anyon system is an interesting theoretical model in its own right. The superconducting behavior of anyon systems at $T=0$ has been investigated by many authors \cite{6}-\cite{15a}. Crucial to the existence of anyon superconductivity at $T=0$ is the exact cancellation between the bare and induced Chern-Simons terms in the effective action of the theory. Although a general consensus exists regarding the superconductivity of anyon systems at zero temperature, a similar consensus at finite temperature is yet to be achieved \cite{8}-\cite{16}. Some authors (see ref. \cite{9}) have concluded that the superconductivity is lost at $T\neq 0$, based upon the appearance of a temperature-dependent correction to the induced Chern-Simons coefficient that is not cancelled out by the bare term. In ref. \cite{11} it is argued, however, that this finite temperature correction is numerically negligible at $T<200$ $K$, and that the main reason for the lack of a Meissner effect is the development of a pole $\sim \left( \frac{1}{{\bf k}% ^{2}}\right) $ in the polarization operator component $\Pi _{00}$ at $T\neq 0 $. There, it is discussed how the existence of this pole leads to a so called partial Meissner effect with a constant magnetic field penetration throughout the sample that appreciably increases with temperature. On the other hand, in ref. \cite{8}, it has been independently claimed that the anyon model cannot superconduct at finite temperature due to the existence of a long-range mode, found inside the infinite bulk at $T\neq 0$. The long range mode found in ref. \cite{8} is also a consequence of the existence of a pole $\sim \left( \frac{1}{{\bf k}^{2}}\right) $ in the polarization operator component $\Pi _{00}$ at $T\neq 0$. The apparent lack of superconductivity at temperatures greater than zero has been considered as a discouraging property of anyon models. Nevertheless, it may be still premature to disregard the anyons as a feasible solution for explaining high -T$_{c}$ superconductivity, at least if the reason sustaining such a belief is the absence of the Meissner effect at finite temperature. As it was shown in a previous paper \cite{16}, the lack of a Meissner effect, reported in ref. \cite{11} for the case of a half-plane sample as a partial Meissner effect, is a direct consequence of the omission of the sample boundary effects in the calculations of the minimal solution for the magnetic field within the sample. To understand this remark we must take into account that the results of ref. \cite{11} were obtained by finding the magnetization in the bulk due to an externally applied magnetic field at the boundary of a half-plane sample. However, in doing so, a uniform magnetization was assumed and therefore the boundary effects were indeed neglected. Besides, in ref. \cite{11} the field equations were solved considering only one short-range mode of propagation for the magnetic field, while as has been emphasized in our previous letter \cite{16}, there is a second short-range mode whose qualitative contribution to the solutions of the field equations cannot be ignored. In the present paper we study the effects of the sample's boundaries in the magnetic response of the anyon fluid at finite temperature. This is done by considering a sample shaped as an infinite strip. When a constant and homogeneous external magnetic field, which is perpendicular to the sample plane, is applied at the boundaries of the strip, two different magnetic responses, depending on the temperature values, can be identified. At temperatures smaller than the fermion energy gap inherent to the many-particle MCS model ($T\ll \omega _{c}$), the system exhibits a Meissner effect. In this case the magnetic field cannot penetrate the bulk farther than a very short distance ($\overline{\lambda }\sim 10^{-5}cm$ for electron densities characteristic of the high -T$_{c}$ superconductors and $T\sim 200$ $K$). On the other hand, as it is natural to expect from a physical point of view, when the temperatures are larger than the energy gap ($T\gg \omega _{c} $) the Meissner effect is lost. In this temperature region a periodic inhomogeneous magnetic field is present within the bulk. These results, together with those previously reported in ref. \cite{16}, indicate that, contrary to some authors' belief, the superconducting behavior (more precisely, the Meissner effect), found in the charged anyon fluid at $T=0$, does not disappear as soon as the system is heated. As it is shown below, the presence of boundaries can affect the dynamics of the system in such a way that the mode that accounts for a homogeneous field penetration \cite{8} cannot propagate in the bulk. Although these results have been proved for two types of samples, the half-plane \cite{16} and the infinite strip reported in this paper, we conjecture that similar effects should also exist in other geometries. Our main conclusion is that the magnetic behavior of the anyon fluid is not just determined by its bulk properties, but it is essentially affected by the sample boundary conditions. The importance of the boundary conditions in 2+1 dimensional models has been previously stressed in ref.\cite{b}. The plan for the paper is as follows. In Sec. 2, for completeness as well as for the convenience of the reader, we define the many-particle 2+1 dimensional MCS model used to describe the charged anyon fluid, and briefly review its main characteristics. In Sec. 3 we study the magnetic response in the self-consistent field approximation of a charged anyon fluid confined to an infinite-strip, finding the analytical solution of the MCS field equations that satisfies the boundary conditions. The fermion contribution in this approximation is given by the corresponding polarization operators at $T\neq 0$ in the background of a many-particle induced Chern-Simons magnetic field. Using these polarization operators in the low temperature approximation ($T\ll \omega _{c}$), we determine the system's two London penetration depths. Taking into account that the boundary conditions are not enough to completely determine the magnetic field solution within the sample, an extra physical condition, the minimization of the system free-energy density, is imposed. This is done in Sec. 4. In this section we prove that even though the electromagnetic field has a long-range mode of propagation in the charged anyon fluid at $T\neq 0$ \cite{8}, a constant and uniform magnetic field applied at the sample's boundaries cannot propagate through this mode. The explicit temperature dependence at $T\ll \omega _{c}$ of all the coefficients appearing in the magnetic field solution, and of the effective London penetration depth are also found. In Sec. 5, we discuss how the superconducting behavior of the charged anyon fluid disappears at temperatures larger than the energy gap ($T\gg \omega _{c}$). Sec. 6 contains the summary and discussion. \section{MCS Many-Particle Model} The Lagrangian density of the 2+1 dimensional non-relativistic charged MCS system is \begin{equation} {\cal L}=-\frac{1}{4}F_{\mu \nu }^{2}-\frac{N}{4\pi }\varepsilon ^{\mu \nu \rho }a_{\mu }\partial _{\nu }a_{\rho }+en_{e}A_{0}+i\psi ^{\dagger }D_{0}\psi -\frac{1}{2m}\left| D_{k}\psi \right| ^{2}+\psi ^{\dagger }\mu \psi \eqnum{2.1} \end{equation} where $A_{\mu }$ and $a_{\mu }$ represent the electromagnetic and the Chern-Simons fields respectively. The role of the Chern-Simons fields is simply to change the quantum statistics of the matter field, thus, they do not have an independent dynamics. $\psi $ represents non-relativistic spinless fermions. $N\ $ is a positive integer that determines the magnitude of the Chern-$% \mathop{\rm Si}% $mons coupling constant. The charged character of the system is implemented by introducing a chemical potential $\mu $; $n_{e}$ is a background neutralizing `classical' charge density, and $m$ is the fermion mass. We will consider throughout the paper the metric $g_{\mu \nu }$=$(1,-% \overrightarrow{1})$. The covariant derivative $D_{\nu }$ is given by \begin{equation} D_{\nu }=\partial _{\nu }+i\left( a_{\nu }+eA_{\nu }\right) ,\qquad \nu =0,1,2 \eqnum{2.2} \end{equation} It is known that to guarantee the system neutrality in the presence of a different from zero fermion density $\left( n_{e}\neq 0\right) $,$\ $a nontrivial background of Chern-Simons magnetic field $\left( \overline{b}=% \overline{f}_{21}\right) $ is generated. The Chern-Simons background field is obtained as the solution of the mean field Euler-Lagrange equations derived from (2.1) \begin{mathletters} \begin{equation} -\frac{N}{4\pi }\varepsilon ^{\mu \nu \rho }f_{\nu \rho }=\left\langle j^{\mu }\right\rangle \eqnum{2.3} \end{equation} \begin{equation} \partial _{\nu }F^{\mu \nu }=e\left\langle j^{\mu }\right\rangle -en_{e}\delta ^{\mu 0} \eqnum{2.4} \end{equation} considering that the system formed by the electron fluid and the background charge $n_{e}$ is neutral \end{mathletters} \begin{equation} \left\langle j^{0}\right\rangle -n_{e}\delta ^{\mu 0}=0 \eqnum{2.5} \end{equation} In eq. (2.5) $\left\langle j^{0}\right\rangle $ is the fermion density of the many-particle fermion system \begin{equation} \left\langle j^{0}\right\rangle =\frac{\partial \Omega }{\partial \mu }, \eqnum{2.6} \end{equation} $\Omega $ is the fermion thermodynamic potential. In this approximation it is found from (2.3)-(2.5) that the Chern-Simons magnetic background is given by \begin{equation} \overline{b}=\frac{2\pi n_{e}}{N} \eqnum{2.7} \end{equation} Then, the unperturbed one-particle Hamiltonian of the matter field represents a particle in the background of the Chern-Simons magnetic field $% \overline{b\text{,}}$ \begin{equation} H_{0}=-\frac{1}{2m}\left[ \left( \partial _{1}+i\overline{b}x_{2}\right) ^{2}+\partial _{2}^{2}\right] \eqnum{2.8} \end{equation} In (2.8) we considered the background Chern-Simons potential, $\overline{a}% _{k}$, $(k=1,2)$, in the Landau gauge \begin{equation} \overline{a}_{k}=\overline{b}x_{2}\delta _{k1} \eqnum{2.9} \end{equation} The eigenvalue problem defined by the Hamiltonian (2.8) with periodic boundary conditions in the $x_{1}$-direction: $\Psi \left( x_{1}+L,\text{ }% x_{2}\right) =$ $\Psi \left( x_{1},\text{ }x_{2}\right) $, \begin{equation} H_{0}\Psi _{nk}=\epsilon _{n}\Psi _{nk},\qquad n=0,1,2,...\text{ }and\text{ }% k\in {\cal Z} \eqnum{2.10} \end{equation} has eigenvalues and eigenfunctions given respectively by \begin{equation} \epsilon _{n}=\left( n+\frac{1}{2}\right) \omega _{c}\qquad \eqnum{2.11} \end{equation} \begin{equation} \Psi _{nk}=\frac{\overline{b}^{1/4}}{\sqrt{L}}\exp \left( -2\pi ikx_{1}/L\right) \varphi _{n}\left( x_{2}\sqrt{\overline{b}}-\frac{2\pi k}{L% \sqrt{\overline{b}}}\right) \eqnum{2.12} \end{equation} where $\omega _{c}=\overline{b}/m$ is the cyclotron frequency and $\varphi _{n}\left( \xi \right) $ are the orthonormalized harmonic oscillator wave functions. Note that the energy levels $\epsilon _{n}$ are degenerates (they do not depend on $k$). Then, for each Landau level $n$ there exists a band of degenerate states. The cyclotron frequency $\omega _{c}$ plays here the role of the energy gap between occupied Landau levels. It is easy to prove that the filling factor, defined as the ratio between the density of particles $% n_{e}$ and the number of states per unit area of a full Landau level, is equal to the Chern-$% \mathop{\rm Si}% $mons coupling constant $N$. Thus, because we are considering that $N$ is a positive integer, we have in this MCS theory $N$ completely filled Landau levels. Once this ground state is established, it can be argued immediately \cite{6}, \cite{6b}, \cite{10a}, \cite{15}, that at $T=0$ the system will be confined to a filled band, which is separated by an energy gap from the free states; therefore, it is natural to expect that at $T=0$ the system should superconduct. This result is already a well established fact on the basis of Hartree-Fock analysis\cite{6} and Random Phase Approximation \cite{6b},\cite {10a}. The case at $T\neq 0$ is more controversial since thermal fluctuations, occurring in the many-particle system, can produce significant changes. As we will show in this paper, the presence in this theory of a natural scale, the cyclotron frequency $\omega _{c}$, is crucial for the existence of a phase at $T\ll \omega _{c}$, on which the system, when confined to a bounded region, still behaves as a superconductor. The fermion thermal Green's function in the presence of the background Chern-Simons field $\overline{b}$ \begin{equation} G\left( x,x^{\prime }\right) =-\left\langle T_{\tau }\psi \left( x\right) \overline{\psi }\left( x^{\prime }\right) \right\rangle \eqnum{2.13} \end{equation} is obtained by solving the equation \begin{equation} \left( \partial _{\tau }-\frac{1}{2m}\overline{D}_{k}^{2}-\mu \right) G\left( x,x^{\prime }\right) =-\delta _{3}\left( x-x^{\prime }\right) \eqnum{2.14} \end{equation} subject to the requirement of antiperiodicity under the imaginary time translation $\tau \rightarrow \tau +\beta $ ($\beta $ is the inverse absolute temperature). In (2.14) we have introduced the notation \begin{equation} \overline{D}_{k}=\partial _{k}+i\overline{a}_{k} \eqnum{2.15} \end{equation} The Fourier transform of the fermion thermal Green's function (2.13) \begin{equation} G\left( p_{4},{\bf p}\right) =\int\limits_{0}^{\beta }d\tau \int d{\bf x}% G\left( \tau ,{\bf x}\right) e^{i\left( p_{4}\tau -{\bf px}\right) } \eqnum{2.16} \end{equation} can be expressed in terms of the orthonormalized harmonic oscillator wave functions $\varphi _{n}\left( \xi \right) $ as \cite{Efrain} \begin{eqnarray} G\left( p_{4},{\bf p}\right) &=&\int\limits_{0}^{\infty }d\rho \int\limits_{-\infty }^{\infty }dx_{2}\sqrt{\overline{b}}\exp -\left( ip_{2}x_{2}\right) \exp -\left( ip_{4}+\mu -\frac{\overline{b}}{2m}\right) \rho \nonumber \\ &&\sum\limits_{n=0}^{\infty }\varphi _{n}\left( \xi \right) \varphi _{n}\left( \xi ^{\prime }\right) t^{n} \eqnum{2.17} \end{eqnarray} where $t=\exp \frac{\overline{b}}{m}\rho $, $\xi =\frac{p_{1}}{\sqrt{% \overline{b}}}+\frac{x_{2}\sqrt{\overline{b}}}{2}$, $\xi ^{\prime }=\frac{% p_{1}}{\sqrt{\overline{b}}}-\frac{x_{2}\sqrt{\overline{b}}}{2}$ and $% p_{4}=(2n+1)\pi /\beta $ are the discrete frequencies $(n=0,1,2,...)$ corresponding to fermion fields. \section{Linear Response in the Infinite Strip} \subsection{Effective Theory at $\mu \neq 0$ and $T\neq 0$} In ref.\cite{8} the effective current-current interaction of the MCS model was calculated to determine the independent components of the magnetic interaction at finite temperature in a sample without boundaries, i.e., in the free space. These authors concluded that the pure Meissner effect observed at zero temperature is certainly compromised by the appearance of a long-range mode at $T\neq 0$. Our main goal in the present paper is to investigate the magnetic response of the charged anyon fluid at finite temperature for a sample that confines the fluid within some specific boundaries. As we prove henceforth, the confinement of the system to a bounded region (a condition which is closer to the experimental situation than the free-space case) is crucial for the realization of the Meissner effect inside the charged anyon fluid at finite temperature. Let us investigate the linear response of a charged anyon fluid at finite temperature and density to an externally applied magnetic field in the specific case of an infinite-strip sample. The linear response of the medium can be found under the assumption that the quantum fluctuations of the gauge fields about the ground-state are small. In this case the one-loop fermion contribution to the effective action, obtained after integrating out the fermion fields, can be evaluated up to second order in the gauge fields. The effective action of the theory within this linear approximation \cite{8},% \cite{11} takes the form \begin{equation} \Gamma _{eff}\,\left( A_{\nu },a_{\nu }\right) =\int dx\left( -\frac{1}{4}% F_{\mu \nu }^{2}-\frac{N}{4\pi }\varepsilon ^{\mu \nu \rho }a_{\mu }\partial _{\nu }a_{\rho }+en_{e}A_{0}\right) +\Gamma ^{\left( 2\right) } \eqnum{3.1} \end{equation} \[ \Gamma ^{\left( 2\right) }=\int dx\Pi ^{\nu }\left( x\right) \left[ a_{\nu }\left( x\right) +eA_{\nu }\left( x\right) \right] +\int dxdy\left[ a_{\nu }\left( x\right) +eA_{\nu }\left( x\right) \right] \Pi ^{\mu \nu }\left( x,y\right) \left[ a_{\nu }\left( y\right) +eA_{\nu }\left( y\right) \right] \] Here $\Gamma ^{\left( 2\right) }$ is the one-loop fermion contribution to the effective action in the linear approximation. The operators $\Pi _{\nu }$ and $\Pi _{\mu \nu }$ are calculated using the fermion thermal Green's function in the presence of the background field $\overline{b}$ (2.17). They represent the fermion tadpole and one-loop polarization operators respectively. Their leading behaviors for static $\left( k_{0}=0\right) $ and slowly $\left( {\bf k}\sim 0\right) $ varying configurations in the frame ${\bf k}=(k,0)$ take the form \begin{equation} \Pi _{k}\left( x\right) =0,\;\;\;\Pi _{0}\left( x\right) =-n_{e},\;\;\;\Pi _{\mu \nu }=\left( \begin{array}{ccc} {\it \Pi }_{{\it 0}}+{\it \Pi }_{{\it 0}}\,^{\prime }\,k^{2} & 0 & {\it \Pi }% _{{\it 1}}k \\ 0 & 0 & 0 \\ -{\it \Pi }_{{\it 1}}k & 0 & {\it \Pi }_{\,{\it 2}}k^{2} \end{array} \right) , \eqnum{3.2} \end{equation} The independent coefficients: ${\it \Pi }_{{\it 0}}$, ${\it \Pi }_{{\it 0}% }\,^{\prime }$, ${\it \Pi }_{{\it 1}}$ and ${\it \Pi }_{\,{\it 2}}$ are functions of $k^{2}$, $\mu $ and $\overline{b}$. In order to find them we just need to calculate the $\Pi _{\mu \nu }$ Euclidean components: $\Pi _{44} $, $\Pi _{42}$ and $\Pi _{22}$. In the Landau gauge these Euclidean components are given by\cite{11}, \begin{mathletters} \begin{equation} \Pi _{44}\left( k,\mu ,\overline{b}\right) =-\frac{1}{\beta }% \sum\limits_{p_{4}}\frac{d{\bf p}}{\left( 2\pi \right) ^{2}}G\left( p\right) G\left( p-k\right) , \eqnum{3.3} \end{equation} \begin{equation} \Pi _{4j}\left( k,\mu ,\overline{b}\right) =\frac{i}{2m\beta }% \sum\limits_{p_{4}}\frac{d{\bf p}}{\left( 2\pi \right) ^{2}}\left\{ G\left( p\right) \cdot D_{j}^{-}G\left( p-k\right) +D_{j}^{+}G\left( p\right) \cdot G\left( p-k\right) \right\} , \eqnum{3.4} \end{equation} \end{mathletters} \begin{eqnarray} \Pi _{jk}\left( k,\mu ,\overline{b}\right) &=&\frac{1}{4m^{2}\beta }% \sum\limits_{p_{4}}\frac{d{\bf p}}{\left( 2\pi \right) ^{2}}\left\{ D_{k}^{-}G\left( p\right) \cdot D_{j}^{-}G\left( p-k\right) +D_{j}^{+}G\left( p\right) \cdot D_{k}^{+}G\left( p-k\right) \right. \nonumber \\ &&\left. +D_{j}^{+}D_{k}^{-}G\left( p\right) \cdot G\left( p-k\right) +G\left( p\right) \cdot D_{j}^{-}D_{k}^{+}G\left( p-k\right) \right\} \nonumber \\ &&-\frac{1}{2m}\Pi _{4}, \eqnum{3.5} \end{eqnarray} where the notation \begin{eqnarray} D_{j}^{\pm }G\left( p\right) &=&\left[ ip_{j}\mp \frac{\overline{b}}{2}% \varepsilon ^{jk}\partial _{p_{k}}\right] G\left( p\right) , \nonumber \\ D_{j}^{\pm }G\left( p-k\right) &=&\left[ i\left( p_{j}-k_{j}\right) \mp \frac{\overline{b}}{2}\varepsilon ^{jk}\partial _{p_{k}}\right] G\left( p-k\right) , \eqnum{3.6} \end{eqnarray} was used. Using (3.3)-(3.5) after summing in $p_{4}$, we found that, in the $k/\sqrt{% \overline{b}}\ll 1$ limit, the polarization operator coefficients ${\it \Pi }% _{{\it 0}}$, ${\it \Pi }_{{\it 0}}\,^{\prime }$, ${\it \Pi }_{{\it 1}}$ and $% {\it \Pi }_{\,{\it 2}}$ are \[ {\it \Pi }_{{\it 0}}=\frac{\beta \overline{b}}{8\pi {\bf k}^{2}}% \sum_{n}\Theta _{n},\;\qquad {\it \Pi }_{{\it 0}}\,^{\prime }=\frac{2m}{\pi \overline{b}}\sum_{n}\Delta _{n}-\frac{\beta }{8\pi }\sum_{n}(2n+1)\Theta _{n}, \] \[ {\it \Pi }_{{\it 1}}=\frac{1}{\pi }\sum_{n}\Delta _{n}-\frac{\beta \overline{% b}}{16\pi m}\sum_{n}(2n+1)\Theta _{n},\qquad {\it \Pi }_{\,{\it 2}}=\frac{1}{% \pi m}\sum_{n}(2n+1)\Delta _{n}-\frac{\beta \overline{b}}{32\pi m^{2}}% \sum_{n}(2n+1)^{2}\Theta _{n}, \] \begin{equation} \Theta _{n}=% \mathop{\rm sech}% \,^{2}\frac{\beta (\epsilon _{n}/2-\mu )}{2},\qquad \Delta _{n}=(e^{\beta (\epsilon _{n}/2-\mu )}+1)^{-1} \eqnum{3.7} \end{equation} The leading contributions of the one-loop polarization operator coefficients (3.7) at low temperatures $\left( T\ll \omega _{c}\right) $ are \begin{equation} {\it \Pi }_{{\it 0}}=\frac{2\beta \overline{b}}{\pi }e^{-\beta \overline{b}% /2m},\qquad {\it \Pi }_{{\it 0}}\,^{\prime }=\frac{mN}{2\pi \overline{b}}% {\it \Lambda },\qquad {\it \Pi }_{{\it 1}}=\frac{N}{2\pi }{\it \Lambda }% ,\quad {\it \Pi }_{\,{\it 2}}=\frac{N^{2}}{4\pi m}{\it \Lambda },\qquad {\it % \Lambda }=\left[ 1-\frac{2\beta \overline{b}}{m}e^{-\beta \overline{b}% /2m}\right] \eqnum{3.8} \end{equation} and at high temperatures $\left( T\gg \omega _{c}\right) $ are \begin{equation} {\it \Pi }_{{\it 0}}=\frac{m}{2\pi }\left[ \tanh \frac{\beta \mu }{2}% +1\right] ,\qquad {\it \Pi }_{{\it 0}}\,^{\prime }=-\frac{\beta }{48\pi }% \mathop{\rm sech}% \!^{2}\!\,\left( \frac{\beta \mu }{2}\right) ,\qquad {\it \Pi }_{{\it 1}}=% \frac{\overline{b}}{m}{\it \Pi }_{{\it 0}}\,^{\prime },\qquad {\it \Pi }_{\,% {\it 2}}=\frac{1}{12m^{2}}{\it \Pi }_{{\it 0}} \eqnum{3.9} \end{equation} In these expressions $\mu $ is the chemical potential and $m=2m_{e}$ ($m_{e}$ is the electron mass). These results are in agreement with those found in refs.\cite{8},\cite{14}. \subsection{MCS Linear Equations} To find the response of the anyon fluid to an externally applied magnetic field, one needs to use the extremum equations derived from the effective action (3.1). This formulation is known in the literature as the self-consistent field approximation\cite{11}. In solving these equations we confine our analysis to gauge field configurations which are static and uniform in the y-direction. Within this restriction we take a gauge in which $A_{1}=a_{1}=0$. The Maxwell and Chern-Simons extremum equations are respectively, \begin{equation} \partial _{\nu }F^{\nu \mu }=eJ_{ind}^{\mu } \eqnum{3.10a} \end{equation} \begin{equation} -\frac{N}{4\pi }\varepsilon ^{\mu \nu \rho }f_{\nu \rho }=J_{ind}^{\mu } \eqnum{3.10b} \end{equation} Here, $f_{\mu \nu }$ is the Chern-Simons gauge field strength tensor, defined as $f_{\mu \nu }=\partial _{\mu }a_{\nu }-\partial _{\nu }a_{\mu }$, and $J_{ind}^{\mu }$ is the current density induced by the anyon system at finite temperature and density. Their different components are given by \begin{equation} J_{ind}^{0}\left( x\right) ={\it \Pi }_{{\it 0}}\left[ a_{0}\left( x\right) +eA_{0}\left( x\right) \right] +{\it \Pi }_{{\it 0}}\,^{\prime }\partial _{x}\left( {\cal E}+eE\right) +{\it \Pi }_{{\it 1}}\left( b+eB\right) \eqnum{3.11a} \end{equation} \begin{equation} J_{ind}^{1}\left( x\right) =0,\qquad J_{ind}^{2}\left( x\right) ={\it \Pi }_{% {\it 1}}\left( {\cal E}+eE\right) +{\it \Pi }_{\,{\it 2}}\partial _{x}\left( b+eB\right) \eqnum{3.11b} \end{equation} in the above expressions we used the following notation: ${\cal E}=f_{01}$, $% E=F_{01}$, $b=f_{12}$ and $B=F_{12}$. Eqs. (3.11) play the role in the anyon fluid of the London equations in BCS superconductivity. When the induced currents (3.11) are substituted in eqs. (3.10) we find, after some manipulation, the set of independent differential equations, \begin{equation} \omega \partial _{x}^{2}B+\alpha B=\gamma \left[ \partial _{x}E-\sigma A_{0}\right] +\tau \,a_{0}, \eqnum{3.12} \end{equation} \begin{equation} \partial _{x}B=\kappa \partial _{x}^{2}E+\eta E, \eqnum{3.13} \end{equation} \begin{equation} \partial _{x}a_{0}=\chi \partial _{x}B \eqnum{3.14} \end{equation} The coefficients appearing in these differential equations depend on the components of the polarization operators through the relations \[ \omega =\frac{2\pi }{N}{\it \Pi }_{{\it 0}}\,^{\prime },\quad \alpha =-e^{2}% {\it \Pi }_{{\it 1}},\quad \tau =e{\it \Pi }_{{\it 0}},\quad \chi =\frac{% 2\pi }{eN},\quad \sigma =-\frac{e^{2}}{\gamma }{\it \Pi }_{{\it 0}},\quad \eta =-\frac{e^{2}}{\delta }{\it \Pi }_{{\it 1}}, \] \begin{equation} \gamma =1+e^{2}{\it \Pi }_{{\it 0}}\,^{\prime }-\frac{2\pi }{N}{\it \Pi }_{% {\it 1}},\quad \delta =1+e^{2}{\it \Pi }_{\,{\it 2}}-\frac{2\pi }{N}{\it \Pi }_{{\it 1}},\quad \kappa =\frac{2\pi }{N\delta }{\it \Pi }_{\,{\it 2}}. \eqnum{3.15} \end{equation} Distinctive of eq. (3.12) is the presence of the nonzero coefficients $% \sigma $ and $\tau $. They are related to the Debye screening in the two dimensional anyon thermal ensemble. A characteristic of this 2+1 dimensional model is that the Debye screening disappears at $T=0$, even if the chemical potential is different from zero. Note that $\sigma $ and $\tau $ link the magnetic field to the zero components of the gauge potentials, $A_{0}$ and $% a_{0}$. As a consequence, these gauge potentials will play a nontrivial role in finding the magnetic field solution of the system. \subsection{Field Solutions and Boundary Conditions} Using eqs.(3.12)-(3.14) one can obtain a higher order differential equation that involves only the electric field, \begin{equation} a\partial _{x}^{4}E+d\partial _{x}^{2}E+cE=0, \eqnum{3.16} \end{equation} Here, $a=\omega \kappa $, $d=\omega \eta +\alpha \kappa -\gamma -\tau \kappa \chi $, and $c=\alpha \eta -\sigma \gamma -\tau \eta \chi $. Solving (3.16) we find \begin{equation} E\left( x\right) =C_{1}e^{-x\xi _{1}}+C_{2}e^{x\xi _{1}}+C_{3}e^{-x\xi _{2}}+C_{4}e^{x\xi _{2}}, \eqnum{3.17} \end{equation} where \begin{equation} \xi _{1,2}=\left[ -d\pm \sqrt{d^{2}-4ac}\right] ^{\frac{1}{2}}/\sqrt{2a} \eqnum{3.18} \end{equation} When the low density approximation $n_{e}\ll m^{2}$ is considered (this approximation is in agreement with the typical values $n_{e}=2\times 10^{14}cm^{-2}$, $m_{e}=2.6\times 10^{10}cm^{-1}$ found in high-T$_{C}$ superconductivity), we find that, for $N=2$ and at temperatures lower than the energy gap $\left( T\ll \omega _{c}\right) $, the inverse length scales (3.18) are given by the following real functions \begin{equation} \xi _{1}\simeq e\sqrt{\frac{m}{\pi }}\left[ 1+\left( \frac{\pi ^{2}n_{e}^{2}% }{m^{3}}\right) \beta \exp -\left( \frac{\pi n_{e}\beta }{2m}\right) \right] \eqnum{3.19} \end{equation} \begin{equation} \xi _{2}\simeq e\sqrt{\frac{n_{e}}{m}}\left[ 1-\left( \frac{\pi n_{e}}{m}% \right) \beta \exp -\left( \frac{\pi n_{e}\beta }{2m}\right) \right] \eqnum{3.20} \end{equation} These two inverse length scales correspond to two short-range electromagnetic modes of propagation. These results are consistent with those obtained in ref. \cite{8} using a different approach. If the masses of the two massive modes, obtained in ref. \cite{8} by using the electromagnetic thermal Green's function for static and slowly varying configurations, are evaluated in the range of parameters considered above, it can be shown that they reduce to eqs. (319), (3.20). The solutions for $B$, $a_{0}$ and $A_{0}$, can be obtained using eqs. (3.13), (3.14), (3.17) and the definition of $E$ in terms of $A_{0,}$ \begin{equation} B\left( x\right) =\gamma _{1}\left( C_{2}e^{x\xi _{1}}-C_{1}e^{-x\xi _{1}}\right) +\gamma _{2}\left( C_{4}e^{x\xi _{2}}-C_{3}e^{-x\xi _{2}}\right) +C_{5} \eqnum{3.21} \end{equation} \begin{equation} a_{0}\left( x\right) =\chi \gamma _{1}\left( C_{2}e^{x\xi _{1}}-C_{1}e^{-x\xi _{1}}\right) +\chi \gamma _{2}\left( C_{4}e^{x\xi _{2}}-C_{3}e^{-x\xi _{2}}\right) +C_{6} \eqnum{3.22} \end{equation} \begin{equation} A_{0}\left( x\right) =\frac{1}{\xi _{1}}\left( C_{1}e^{-x\xi _{1}}-C_{2}e^{x\xi _{1}}\right) +\frac{1}{\xi _{2}}\left( C_{3}e^{-x\xi _{2}}-C_{4}e^{x\xi _{2}}\right) +C_{7} \eqnum{3.23} \end{equation} In the above formulas we introduced the notation $\gamma _{1}=\left( \xi _{1}^{2}\kappa +\eta \right) /\xi _{1}$, $\gamma _{2}=\left( \xi _{2}^{2}\kappa +\eta \right) /\xi _{2}$. In obtaining eq. (3.16) we have taken the derivative of eq. (3.12). Therefore, the solution of eq. (3.16) belongs to a wider class than the one corresponding to eqs. (3.12)-(3.14). To exclude redundant solutions we must require that they satisfy eq. (3.12) as a supplementary condition. In this way the number of independent unknown coefficients is reduced to six, which is the number corresponding to the original system (3.12)-(3.14). The extra unknown coefficient is eliminated substituting the solutions (3.17), (3.21), (3.22) and (3.23) into eq. (3.12) to obtain the relation \begin{equation} e{\it \Pi }_{{\it 1}}C_{5}=-{\it \Pi }_{{\it 0}}\left( C_{6}+eC_{7}\right) \eqnum{3.24} \end{equation} Eq. (3.24) has an important meaning, it establishes a connection between the coefficients of the long-range modes of the zero components of the gauge potentials $(C_{6}+eC_{7})$ and the coefficient of the long-range mode of the magnetic field $C_{5}$. Note that if the induced Chern-Simons coefficient ${\it \Pi }_{{\it 1}}$, or the Debye screening coefficient ${\it % \Pi }_{{\it 0}}$ were zero, there would be no link between $C_{5}$ and $% (C_{6}+eC_{7})$. This relation between the long-range modes of $B$, $A_{0}$ and $a_{0}$ can be interpreted as a sort of Aharonov-Bohm effect, which occurs in this system at finite temperature. At $T=0$, we have ${\it \Pi }_{% {\it 0}}=0$, and the effect disappears. Up to this point no boundary has been taken into account. Therefore, it is easy to understand that the magnetic long-range mode associated with the coefficient $C_{5}$, must be identified with the one found in ref. \cite{8} for the infinite bulk using a different approach. However, as it is shown below, when a constant and uniform magnetic field is perpendicularly applied at the boundaries of a two-dimensional sample, this mode cannot propagate (i.e. $C_{5}\equiv 0$) within the sample. This result is crucial for the existence of Meissner effect in this system. In order to determine the unknown coefficients we need to use the boundary conditions. Henceforth we consider that the anyon fluid is confined to the strip $-\infty <y<\infty $ with boundaries at $x=-L$ and $x=L$. The external magnetic field will be applied from the vacuum at both boundaries ($-\infty <x\leq -L$, $\;L\leq x<\infty $). The boundary conditions for the magnetic field are $B\left( x=-L\right) =B\left( x=L\right) =\overline{B}$ ($\overline{B}$ constant). Because no external electric field is applied, the boundary conditions for this field are, $E\left( x=-L\right) =E\left( x=L\right) =0$. Using them and assuming $% L\gg \lambda _{1}$, $\lambda _{2}$ ($\lambda _{1}=1/\xi _{1}$, $\lambda _{2}=1/\xi _{2}$), we find the following relations that give $C_{1,2,3,4}$ in terms of $C_{5}$, \begin{equation} C_{1}=Ce^{-L\xi _{1}},\quad C_{2}=-C_{1},\quad C_{3}=-Ce^{-L\xi _{2}},\quad C_{4}=-C_{3},\quad C=\frac{C_{5}-\overline{B}}{\gamma _{1}-\gamma _{2}} \eqnum{3.25} \end{equation} \section{Stability Condition for the Infinite-Strip Sample} After using the boundary conditions, we can see from (3.25) that they were not sufficient to find the coefficient $C_{5}$. In order to totally determine the system magnetic response we have to use another physical condition from where $C_{5}$ can be found. Since, obviously, any meaningful solution have to be stable, the natural additional condition to be considered is the stability equation derived from the system free energy. With this goal in mind we start from the free energy of the infinite-strip sample \[ {\cal F}=\frac{1}{2}\int\limits_{-L^{\prime }}^{L^{\prime }}dy\int\limits_{-L}^{L}dx\left\{ \left( E^{2}+B^{2}\right) +\frac{N}{\pi }% a_{0}b-{\it \Pi }_{{\it 0}}\left( eA_{0}+a_{0}\right) ^{2}\right. \] \begin{equation} \left. -{\it \Pi }_{{\it 0}}\,^{\prime }\left( eE+{\cal E}\right) ^{2}-2{\it % \Pi }_{{\it 1}}\left( eA_{0}+a_{0}\right) \left( eB+b\right) +{\it \Pi }_{\,% {\it 2}}\left( eB+b\right) ^{2}\right\} \eqnum{4.1} \end{equation} where $L$ and $L^{\prime }$ determine the two sample's lengths. Using the field solutions (3.17), (3.21)-(3.23) with coefficients (3.25), it is found that the leading contribution to the free-energy density ${\it f}=% \frac{{\cal F}}{{\cal A}}$ ,\ (${\cal A}=4LL^{\prime }$ being the sample area) in the infinite-strip limit $(L\gg \lambda _{1}$, $\lambda _{2}$, $% L^{\prime }\rightarrow \infty )$ is given by \begin{equation} f=C_{5}^{2}-{\it \Pi }_{{\it 0}}\left( C_{6}+eC_{7}\right) ^{2}+e^{2}{\it % \Pi }_{\,{\it 2}}C_{5}^{2}-2e{\it \Pi }_{{\it 1}}\left( C_{6}+eC_{7}\right) C_{5} \eqnum{4.2} \end{equation} Taking into account the constraint equation (3.24), the free-energy density (4.2) can be written as a quadratic function in $C_{5}$. Then, the value of $% C_{5}$ is found, by minimizing the corresponding free-energy density \begin{equation} \frac{\delta {\it f}}{\delta C_{5}}=\left[ {\it \Pi }_{{\it 0}}+e^{2}{\it % \Pi }_{{\it 1}}^{\,}\,^{2}+e^{2}{\it \Pi }_{{\it 0}}{\it \Pi }_{\,{\it 2}% }\right] \frac{C_{5}}{{\it \Pi }_{{\it 0}}}=0, \eqnum{4.3} \end{equation} to be $C_{5}=0$. This result implies that the long-range mode cannot propagate within the infinite-strip when a uniform and constant magnetic field is perpendicularly applied at the sample's boundaries. We want to point out the following fact. The same property of the finite temperature polarization operator component $\Pi _{00}$ that is producing the appearance of a long-range mode in the infinite bulk, is also responsible, when it is combined with the boundary conditions, for the non-propagation of this mode in the bounded sample. It is known that the nonvanishing of ${\it \Pi }_{{\it 0}}$ at $T\neq 0$ (or equivalently, the presence of a pole $\sim 1/k^{2}$ in $\Pi _{00}$ at $T\neq 0$) guarantees the existence of a long-range mode in the infinite bulk \cite{8}. On the other hand, however, once ${\it \Pi }_{{\it 0}}$ is different from zero, we can use the constraint (3.24) to eliminate $C_{6}+eC_{7}$ in favor of $C_{5% \text{ }}$ in the free-energy density of the infinite strip. Then, as we have just proved, the only stable solution of this boundary-value problem, which is in agreement with the boundary conditions, is $C_{5}=0$. Consequently, no long-range mode propagates in the bounded sample. In the zero temperature limit $\left( \beta \rightarrow \infty \right) $, because ${\it \Pi }_{{\it 0}}=0$, it is straightforwardly obtained from (3.24) that $C_{5}=0$ and no long-range mode propagates. At $T\neq 0$, taking into account that $C_{5}=0$ along with eq. (3.25) in the magnetic field solution (3.21), we can write the magnetic field penetration as \begin{equation} B\left( x\right) =\overline{B}_{1}\left( T\right) \left( e^{-(x+L)\xi _{1}}+e^{\left( x-L\right) \xi _{1}}\right) +\overline{B}_{2}\left( T\right) \left( e^{-(x+L)\xi _{2}}+e^{\left( x-L\right) \xi _{2}}\right) \eqnum{4.4} \end{equation} where, \begin{equation} \overline{B}_{1}\left( T\right) =\frac{\gamma _{1}}{\gamma _{1}-\gamma _{2}}% \overline{B},\text{ \qquad \quad }\overline{B}_{2}\left( T\right) =\frac{% \gamma _{2}}{\gamma _{2}-\gamma _{1}}\overline{B} \eqnum{4.5} \end{equation} For densities $n_{e}\ll m^{2}$, the coefficients $\overline{B}_{1}$and $% \overline{B}_{2}$ can be expressed, in the low temperature approximation $% \left( T\ll \omega _{c}\right) $, as \begin{equation} B_{1}\left( T\right) \simeq -\frac{\left( \pi n_{e}\right) ^{3/2}}{m^{2}}% \left[ 1/m+\frac{5}{2}\beta \exp -\left( \frac{\pi n_{e}\beta }{2m}\right) \right] \overline{B},\qquad \eqnum{4.6} \end{equation} \begin{equation} B_{2}\left( T\right) \simeq \left[ 1+\frac{5\pi n_{e}}{2m^{2}}\sqrt{\pi n_{e}% }\beta \exp -\left( \frac{\pi n_{e}\beta }{2m}\right) \right] \overline{B} \eqnum{4.7} \end{equation} Hence, in the infinite-strip sample the applied magnetic field is totally screened within the anyon fluid on two different scales, $\lambda _{1}=1/\xi _{1}$ and $\lambda _{2}=1/\xi _{2}$. At $T=200K$, for the density value considered above, the penetration lengths are given by $\lambda _{1}\simeq 0.6\times 10^{-8}cm$ and $\lambda _{2}\simeq 0.1\times 10^{-4}cm$ . Moreover, taking into account that $\xi _{1}$ increases with the temperature while $\xi _{2}$ decreases (see eqs. (3.19)-(3.20)), and that $B_{1}\left( T\right) <0$ while $B_{2}\left( T\right) >0$, it can be shown that the effective penetration length $\overline{\lambda }$ (defined as the distance $% x$ where the magnetic field falls down to a value $B\left( \overline{\lambda }\right) /\overline{B}=e^{-1}$) increases with the temperature as \begin{equation} \overline{\lambda }\simeq \overline{\lambda }_{0}\left( 1+\overline{\kappa }% \beta \exp -\frac{1}{2}\overline{\kappa }\beta \right) \eqnum{4.8} \end{equation} where $\overline{\lambda }_{0}=\sqrt{m/n_{e}e^{2}}$ and $\overline{\kappa }% =\pi n_{e}/m$. At $T=200K$ the effective penetration length is $\overline{% \lambda }\sim 10^{-5}cm$. It is timely to note that the presence of explicit (proportional to $N$) and induced (proportional to ${\it \Pi }_{{\it 1}}$) Chern-Simons terms in the anyon effective action (3.1) is crucial to obtain the Meissner solution (4.4). If the Chern-Simons interaction is disconnected ($N\rightarrow \infty $ and ${\it \Pi }_{{\it 1}}=0$), then $a=0,$ $d=1+e^{2}{\it \Pi }_{{\it 0}% }{}^{\prime }\neq 0$ and $c=e^{2}{\it \Pi }_{{\it 0}}\,\neq 0$ in eq. (3.16). In that case the solution of the field equations within the sample are $E=0$, $B=\overline{B}$. That is, we regain the QED in (2+1)-dimensions, which does not exhibit any superconducting behavior. \section{High Temperature Non-Superconducting Phase} We have just found that the charged anyon fluid confined to an infinite strip exhibits Meissner effect at temperatures lower than the energy gap $% \omega _{c}$. It is natural to expect that at temperatures larger than the energy gap this superconducting behavior should not exist. At those temperatures the electron thermal fluctuations should make available the free states existing beyond the energy gap. As a consequence, the charged anyon fluid should not be a perfect conductor at $T\gg \omega _{c}$. A signal of such a transition can be found studying the magnetic response of the system at those temperatures. As can be seen from the magnetic field solution (4.4), the real character of the inverse length scales (3.18) is crucial for the realization of the Meissner effect. At temperatures much lower than the energy gap this is indeed the case, as can be seen from eqs. (3.19) and (3.20). In the high temperature $\left( T\gg \omega _{c}\right) $ region the polarization operator coefficients are given by eq. (3.9). Using this approximation together with the assumption $n_{e}\ll m^{2}$, we can calculate the coefficients $a$, $c$ and $d$ that define the behavior of the inverse length scales, \begin{equation} a\simeq \pi ^{2}{\it \Pi }_{{\it 0}}{}^{\prime }{\it \Pi }_{\,{\it 2}} \eqnum{5.1} \end{equation} \begin{equation} c\simeq e^{2}{\it \Pi }_{{\it 0}}{} \eqnum{5.2} \end{equation} \begin{equation} d\simeq -1 \eqnum{5.3} \end{equation} Substituting with (5.1)-(5-3) in eq. (3.18) we obtain that the inverse length scales in the high-temperature limit are given by \begin{equation} \xi _{1}\simeq e\sqrt{m/2\pi }\left( \tanh \frac{\beta \mu }{2}+1\right) ^{% \frac{1}{2}} \eqnum{5.4} \end{equation} \begin{equation} \xi _{2}\simeq i\left[ 24\sqrt{\frac{2m}{\beta }}\cosh \frac{\beta \mu }{2}% \left( \tanh \frac{\beta \mu }{2}+1\right) ^{-\frac{1}{2}}\right] \eqnum{5.5} \end{equation} The fact that $\xi _{2}$ becomes imaginary at temperatures larger than the energy gap, $\omega _{c}$, implies that the term $\gamma _{2}\left( C_{4}e^{x\xi _{2}}-C_{3}e^{-x\xi _{2}}\right) $ in the magnetic field solution (3.21) ceases to have a damping behavior, giving rise to a periodic inhomogeneous penetration. Therefore, the fluid does not exhibit a Meissner effect at those temperatures since the magnetic field will not be totally screened. This corroborate our initial hypothesis that at $T\gg \omega _{c}$ the anyon fluid is in a new phase in which the magnetic field can penetrate the sample. We expect that a critical temperature of the order of the energy gap ($T\sim \omega _{c}$) separates the superconducting phase $\left( T\ll \omega _{c}\right) $ from the non-superconducting one $\left( T\gg \omega _{c}\right) $. Nevertheless, the temperature approximations (3.8) and (3.9) are not suitable to perform the calculation needed to find the phase transition temperature. The field solutions in this new non-superconducting phase is currently under investigation. The results will be published elsewhere. \section{Concluding Remarks} In this paper we have investigated the magnetic response at finite temperature of a charged anyon fluid confined to an infinite strip. The charged anyon fluid was modeled by a (2+1)-dimensional MCS theory in a many-particle ($\mu \neq 0$, $\overline{b}\neq 0$) ground state. The particle energy spectrum of the theory exhibits a band structure given by different Landau levels separated by an energy gap $\omega _{c}$, which is proportional to the background Chern-Simons magnetic field $\overline{b}$. We found that the energy gap $\omega _{c}$ defines a scale that separates two phases: a superconducting phase at $T\ll \omega _{c}$, and a non-superconducting one at $T\gg \omega _{c}$. The total magnetic screening in the superconducting phase is characterized by two penetration lengths corresponding to two short-range eigenmodes of propagation of the electromagnetic field within the anyon fluid. The existence of a Meissner effect at finite temperature is the consequence of the fact that a third electromagnetic mode, of a long-range nature, which is present at finite temperature in the infinite bulk \cite{8}, does not propagate within the infinite strip when a uniform and constant magnetic field is applied at the boundaries. This is a significant property since the samples used to test the Meissner effect in high-$T_{c}$ superconductors are bounded. It is noteworthy that the existence at finite temperature of a Debye screening (${\it \Pi }_{{\it 0}}\,\neq 0$) gives rise to a sort of Aharonov-Bohm effect in this system with Chern-Simons interaction ($N$ finite, ${\it \Pi }_{{\it 1}}\neq 0$). When ${\it \Pi }_{{\it 0}}\,\neq 0$, the field combination $a_{0}+eA_{0}$ becomes physical because it enters in the field equations in the same foot as the electric and magnetic fields (see eq. (3.12)). A direct consequence of this fact is that the coefficient $% C_{5}$, associated to the long-range mode of the magnetic field, is linked to the coefficients $C_{6}$ and $C_{7}$ of the zero components of the potentials (see eq. (3.24)). When $T=0$, since ${\it \Pi }_{{\it 0}}\,=0$ and ${\it \Pi }_{{\it 1}}\neq 0$% , eq. (3.24) implies $C_{5}=0$. That is, at zero temperature the long-range mode is absent. This is the well known Meissner effect of the anyon fluid at $T=0$. When $T\neq 0$, eq. (3.24) alone is not enough to determine the value of $C_{5}$, since it is given in terms of $C_{6}$ and $C_{7}$ which are unknown. However, when eq. (3.24) is taken together with the field configurations that satisfy the boundary conditions for the infinite-strip sample (eqs. (3.17), (3.21)-(3.23) and (3.25)), and with the sample stability condition (4.3), we obtain that $C_{5}=0$. Thus, the combined action of the boundary conditions and the Aharonov-Bohm effect expressed by eq. (3.24) accounts for the total screening of the magnetic field in the anyon fluid at finite temperature. Finally, at temperatures large enough ($T\gg \omega _{c}$) to excite the electrons beyond the energy gap, we found that the superconducting behavior of the anyon fluid is lost. This result was achieved studying the nature of the characteristic lengths (3.18) in this high temperature approximation. We showed that in this temperature region the characteristic length $\xi _{2}$ becomes imaginary (eq. (5.5)), which means that a total damping solution for the magnetic field does not exist any more, and hence the magnetic field penetrates the sample. \begin{quote} Acknowledgments \end{quote} The authors are very grateful for stimulating discussions with Profs. G. Baskaran, A. Cabo, E.S. Fradkin, Y. Hosotani and J. Strathdee. We would also like to thank Prof. S. Randjbar-Daemi for kindly bringing the publication of ref. \cite{b} to our attention. Finally, it is a pleasure for us to thank Prof. Yu Lu for his hospitality during our stay at the ICTP, where part of this work was done. This research has been supported in part by the National Science Foundation under Grant No. PHY-9414509.
proofpile-arXiv_065-474
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} An unbroken hidden supersymmetric gauge sector, supporting a gaugino condensate $\vev{\lambda\lambda}\sim (10^{14}\ \mbox{GeV})^3,$ can through its coupling to the dilaton (and other moduli) serve the related dynamical functions of providing a superpotential for the dilaton and a mass for the gravitino \cite{Nil}. Because of its possibly central role in the understanding of these phenomena, the dynamics associated with this hidden gauge sector may merit additional study. As a zero temperature field theory, against a Minkowski space-time background, this has been done (without dilaton) in the original paper of Veneziano and Yankielowicz \cite{Ven} (see also Amati et al \cite{Ama}). In the effective theory, the ground state emerges as containing a gaugino condensate $\vev{\lambda\lambda}$ with residual $Z_N$ symmetry. The changes which evolve with the introduction of supergravity and the dilaton are discussed in \cite{Fer}. In the context of the standard Robertson-Walker cosmology, it is of interest to examine the transition from the hot unconfined phase in the hidden sector to the confined phase.\footnote{It is possible that screening, rather than confinement, characterizes the phase transition \cite{Gro}. The only property to be used in this paper is the change in the degrees of freedom from perturbative non-singlet quanta to gauge-singlet `hadrons' as one passes through the critical temperature.} Is the transition first or second order? If first order, is the transition completed in a manner consistent with the observed smoothness of the universe? Some possible problems associated with a transition described by the evolution of the condensate as a field theoretic order parameter were discussed in \cite{Gol}. However, the results depended critically on the K\"ahler potential of the effective condensate field, and this is certainly not known. Thus, a more phenomenological approach is indicated. In this work, I present a preliminary study of the transition to the confining phase of the unbroken hidden Yang-Mills sector, under a few well-defined assumptions: \begin{itemize} \itm{1}There exists a hot, unconfined phase of the hidden sector. \itm{2}The transition to confining phase is first order, proceeding through the spontaneous nucleation and expansion of critical bubbles of confined phase. \ei Neither of these assumptions is necessarily correct. The aim of the present study is to simply examine the consequences of taking them to be true The principal conclusion of this study, under the stated assumptions, is the following: if the lightest composite particles in the confined phase (`hidden hadrons') are more massive than $\sim 2\ T_c$ (where $T_c$ is the critical temperature), then the requirement of sufficient supercooling in order to complete the phase transition implies a critical bubble size too small to contain enough ({\em i.e.,} $>100$) hidden hadrons to allow a meaningful thermodynamic description. This result is largely traceable to the high value of the transition temperature. The conclusion will be shown to hold as well in the presence of (near) massless vector-like matter fields. Thus a first order transition in a {\em cosmological} context may be unlikely in the hidden sector. \section{Review of Homogeneous Nucleation} In this section, I will briefly sketch the derivation of the relevant formulae, initially following Fuller, Mathews, and Alcock \cite{Ful} in their discussion of the quark-hadron transition, with some modification for ease of application to the present situation. The formation of the confined phase proceeds through the spontaneous nucleation of critical bubbles of radius $R_c,$ at which the free energy difference between confined and deconfined phases (pressures $P_{\rm conf}(T)$ and $P_{\rm deconf}(T),$ respectively) \be \Delta F= -\frac{4\pi}{3}R_c^3\ (P_{\rm conf}-P_{\rm deconf}) + 4\pi\sigmaR_c^2 \lab{delf} \end{equation} has a saddle point at \be R_c(T) = \frac{2\sigma}{P_{\rm conf}(T)-P_{\rm deconf}(T)}\ \ . \lab{rc} \end{equation} Here $\sigma$ is the surface energy density at the interface between the phases. The probability of nucleation of a single bubble is then \be p(T)=CT^4\ e^{-\Delta F/T} \ \ . \lab{p1} \end{equation} The critical temperature $T_c$ is define through the coexistence condition in the infinite volume limit \be P_{\rm conf}(T_c)=P_{\rm deconf}(T_c)\ \ . \lab{coex} \end{equation} If the amount of supercooling $\eta\equiv (T_c-T)/T_c$ is small, then one may expand \be P_{\rm conf}(T)-P_{\rm deconf}(T)\simeq L\eta\ \ , \lab{delp} \end{equation} where \be L=T_c\ \frac{\partial}{\partial T}\left.(P_{\rm deconf}-P_{\rm conf})\right|_{T_c} \lab{lat} \end{equation} is the latent heat released per unit volume during the phase change. Combining these results, we have \be p(T)\simeq CT_c^4\ e^{-A(T)}\ \ , \lab{p2} \end{equation} where \be A(T)=\frac{16\pi}{3}\ \frac{\sigma^3}{L^2T_c\eta^2} \lab{a} \end{equation} is the 3-dimensional critical bubble action. The rapid increase of $p(T)$ with increased supercooling is then explicit. The condition for the complete nucleation of the universe at time $t_f$ is \be \int^{t_f}_{t_c}dt^{\, \prime}\ p(T^{\, \prime}(t^{\, \prime}))\ \frac{4\pi}{3}\ v_s^3\ (t_f-t^{\, \prime})^3=1\ \ , \lab{nuc1} \end{equation} where $v_s\simeq 1/\sqrt{3}$ is the speed of the expanding shock wave of the confined phase into the metastable unconfined phase. For small supercooling, one may expand about $t^{\, \prime}=t_f:$ \cite{Ful} \be \ln p(T)=\ln p(T_f) + \left.\frac{d\ln p}{dT}\right|_{T_f}\ \left.\frac{dT}{dt}\right|_{t_f}\ (t-t_f) \lab{lnp} \end{equation} with \be \left.\frac{dT}{dt}\right|_{t_f}=-T_f\ H(T_f)\simeq -T_c H(T_c)\ \ . \lab{dtdt} \end{equation} Thus \be p(t)=p(t_f)\ e^{-\lambda(t_f-t)} \lab{pt} \end{equation} with $\lambda=2A(T_f) H(T_c)/\eta_f.$ The condition \eqn{nuc1}\ then becomes \be \frac{8\pi v_s^3}{\lambda^4}\ p(T_f) =1\ \ , \lab{nuc2} \end{equation} or \be \frac{8\pi C}{3\sqrt{3}}\ \left(\frac{T_c}{H(T_c)}\frac{\alpha}{2}\right)^4 = A^6\ e^A \lab{nuc3} \end{equation} where \be \alpha\equiv \left(\frac{16\pi}{3}\ \frac{\sigma^3}{L^2T_c}\right)^{\half}\ \ , \lab{alpha} \end{equation} and $A\equiv A(T_f).$ Note that \be \eta_f=\frac{\alpha}{\sqrt{A}} \lab{etaf} \end{equation} and the Hubble constant at $T_c$ is \be H(T_c)=\frac{1}{M_{\rm Pl}}\ \sqrt{\frac{8\pi\rho(T_c)}{3}}\ \ , \lab{htc} \end{equation} with \be \rho(T_c)=\mbox{total energy density at}\ T_c = \cn_{\rm tot}\frac{\pi^2}{30}\ T_c^4\ \ . \lab{rho} \end{equation} Here $\cn_{\rm tot}$ is an effective statistical weight for the degrees of freedom at $T_c.$ If the universe happens to be dominated by a vacuum energy $\rho_0$ at the time of supercooling, then Eq.~\eqn{rho}\ is just a redefinition of $\rho_0$ in terms of $\cn_{\rm tot}$. Combining Eqs.~\eqn{alpha}-\eqn{rho}, the condition \eqn{nuc3} becomes \be A^6\ e^A = \left(\frac{0.45\ C^{1/4}\ \alpha}{\sqrt{\cn_{\rm tot}}}\ \ \frac{M_{\rm Pl}}{T_c}\right)^4\ \ . \lab{nuc4} \end{equation} \section{Example of Pure Non-Supersymmetric \su{3} Yang Mills} As a preliminary for the hidden sector, consider the case of purely gluonic \su{3}$_c$ (no quarks). Then $\cn_{\rm tot}=16,\ T_c\simeq 200$ MeV, and a lattice-based estimate of $\alpha$ is \cite{Iwa} \be \alpha\simeq 0.0065 \pm 0.0015\ \ . \lab{alphx} \end{equation} With $C\sim 1,$ Eq.~\eqn{nuc4}\ yields \be A=124\ \ , \lab{a3} \end{equation} and Eq.~\eqn{etaf}\ gives $\eta_f= 6\times 10^{-4}.$ From \eqn{rc}, \eqn{delp}, \eqn{alpha}, and \eqn{etaf}, the radius of a critical bubble is \be R_c=\frac{2\sigma}{L\eta_f}=\sqrt{\frac{A}{4\pi}}\ \left(\frac{T_c^3}{\sigma}\right)^{1/2}\ T_c^{-1}\ \ . \lab{rcf} \end{equation} {}From the data of \cite{Iwa}, the surface energy density is not quite scale invariant with present statistics; nevertheless, it will suffice for the present application to take \cite{Bro} \be \sigma\simeq 0.025\ T_c^3\ \ , \lab{sigma} \end{equation} which gives \be R_c= 35\ T_c^{-1}\ \ . \lab{rcx} \end{equation} I now come to the point of departure of the present work. With no quarks, the relevant degrees of freedom in the confined phase are glueballs with masses $M_i \ \,\raisebox{-0.13cm}{$\stackrel{\textstyle>}{\textstyle\sim}$}\,\ 1\ \mbox{GeV}\gg T_c,$ and number densities \be n_i(T_c)\simeq T_c^3\ (2S_i+1) \left(\frac{M_i}{2\piT_c}\right)^{3/2}\ e^{-M_i/T_c}\ \ . \lab{ni} \end{equation} Then the number of glueballs in the critical bubble is \be N_c =\sum n_i(T_c)\ \ \tfrac{4}{3}\piR_c^3\ \ . \lab{nc1} \end{equation} Phenomenological studies \cite{Shu} have suggested that a density of states $\tau(M)\propto M^3$ (for each isospin and hypercharge) provides a good fit to the observed hadron spectrum. I adopt this for the glueball spectrum, and normalize to 1 state in the interval $(M_0^2-\thalf\mu^2,M_0^2+\thalf\mu^2),$ where $M_0$ is the mass of the lightest glueball and $\mu^2$ is the inverse of the Regge slope. This gives for the spectral density \be \tau(M)=(2/M_0^2\mu^2)\ M^3\ \ , \lab{tau} \end{equation} and for the number density \begin{eqnarray} n_c&=&\sum n_i(T_c)=2(2\pi)^{-3/2}M_0^{-2}\mu^{-2}\int_{M_0}^{\infty} dM\ M^3\ (M/T_c)^{3/2}\ e^{-M/T_c}\ \ T_c^3 \nonumber\\ &\equiv& 2(2\pi)^{-3/2}\ (T_c/\mu)^2\ f(M_0/T_c)\ \ T_c^3\ \ . \lab{nc2} \end{eqnarray} Combining \eqn{rcx}, (\ref{eq:nc1}), and (\ref{eq:nc2}), one finds \be N_c=22,800\ (T_c/\mu)^2\ f(M_0/T_c)\ \ . \lab{nc3} \end{equation} A thermodynamic description, and hence a first order phase transition, will be {\em cosmologically} viable only for $N_c\gg 1,$ say $N_c>100.$ Once $(\mu/T_c)$ is specified, this will translate via (\ref{eq:nc3}) to a bound on $M_0.$ Since $\mu^2\sim 1 \ \ \mbox{GeV}^2,$ and $T_c\simeq 250\ \mev,$ one finds \be M_0\le 8\ T_c\simeq 2000\ \mbox{MeV}\ \ . \lab{mbound} \end{equation} Before continuing, I will comment briefly on the question of possible temperature dependence of the glueball masses near $T_c.$ A recent theoretical calculation \cite{Sug} in the context of the dual Ginzburg-Landau theory shows that there is some reduction in mass of the gauge singlet QCD monopole at $T_c,$ although the effect is totally dependent on an assumed (and unknown) temperature dependence of the dual Higgs quartic coupling. There is no basis for supposing that the glueball mass goes to zero (or is even much reduced) at $T_c$ in a strongly first order phase transition, and I will simply reinterpret \eqn{mbound} as a bound on an effective glueball mass, with the expectation that it does not differ significantly from the zero temperature mass. I now turn to examine the implication of these ideas when applied to the hidden gauge sector which is relevant to gaugino condensation and SUSY-breaking. \section{The Hidden Sector: Pure SUSY Yang-Mills} There are several significant differences between the hidden sector SUSY-Yang-Mills theory and the non-SUSY SU(3)$_c$ theory just discussed: \begin{itemize} \item The theory is supersymmetric: there are Majorana fermions in the adjoint representation, and only rudimentary lattice results are available for such a theory \cite{Mon}. \item The vacuum properties are totally unlike those in ordinary QCD \cite{Ama}. \item The gauge group is either much larger than SU(3), or is manifest at a Kac-Moody level $k\ge 2,$ in order that strong coupling sets in at a scale $\sim 10^{14}$ GeV. \ei A question which immediately arises as a consequence of these differences is whether the transition is first order. The simplest approach here is to assume that it is first order as a working hypothesis, and examine the consequences. Since the interface energy $\sigma$ and the parameter $\alpha$ are completely unknown for the hidden sector, the nucleation condition \eqn{nuc4}\ must be rewritten in a suitable manner. I assume that the specific entropy of the hadronic phase is much less than that of the gauge phase, so that I take for the latent heat \be L=T_c\ \left.\frac{dP_{\rm deconf}}{dT}\right|_{T=T_c}=4\ \frac{\pi^2}{90}\ \cn_{\rm hidden}\ T_c^4\ \ . \lab{lat1} \end{equation} It is also convenient to set \be \hat\sigma\equiv \sigma/T_c^3\ \ , \lab{sigh} \end{equation} so that the condition \eqn{nuc4}\ becomes \be \left(\frac{A}{\sh}\right)^6\ e^A=\left(\frac{4.8\ C^{1/4}}{\cn_{\rm hidden}\sqrt{\cn_{\rm tot}}}\ \frac{M_{\rm Pl}}{T_c}\right)^4\ \ . \lab{cond} \end{equation} What is $T_c?$ For zero cosmological constant, the gravitino mass is given in terms of the effective superpotential $W$ and Kahler $K$ by \be m_{3/2}=e^{K/2}|W_{eff}|/M^2\ \ , \lab{m32} \end{equation} where $M=M_{\rm Pl}/\sqrt{8\pi}.$ In the effective theory, with a simple gauge group and no matter fields, one obtains after integrating out the gaugino condensate \cite{deC} \be W_{eff}=-\frac{b}{6{\rm e}}\ M_{\rm string}^3\ e^{-3S/2b}\ \ , \lab{weff} \end{equation} where $b=\beta(g)/g^3=3N/16\pi^2$ for \su{N}, Re $S=1/g_{\rm string}^2\simeq 2.0$ at the correct minimum for the dilaton field $S.$ $M_{\rm string}$ sets the scale for the logarithmic term in the condensate superpotential. The \su{N} theory becomes strong $(g^2/4\pi=1)$ at a scale \be \Lambda=M_{\rm string}\ e^{-S/2b}\ \ , \lab{lambda} \end{equation} so that \be m_{3/2}=e^{K/2}\ \frac{b}{6{\rm e}}\ \frac{\Lambda^3}{M^2}\ \ . \lab{m32a} \end{equation} For $m_{3/2}\simeq 10^3\ \ \mbox{GeV},$ one obtains \be \Lambda=e^{-K/6}\ N^{-1/3}\cdot 1.7\times 10^{14}\ \ \mbox{GeV}\ . \lab{lambda1} \end{equation} As a heuristic example, I will choose as the hidden gauge group SU(6), which is consistent with Eqs.~\eqn{lambda}\ and \eqn{lambda1}\ for $M_{\rm string}\simeq 10^{18}\ \ \mbox{GeV}, \ e^{-K/6}\simeq 1.$ With $T_c\simeq \Lambda\simeq 10^{14}\ \ \mbox{GeV},\ \cn_{\rm hidden}=2\left(\frac{15}{8}\right)\left(6^2-1\right)=131.25,$\ $\cn_{\rm tot}= {\cal N}(Standard \ Model)+\cn_{\rm hidden}=213.75+131.25=345,$ and $C^{1/4}\simeq 1,$ the constraint Eq.~\eqn{cond} becomes \be A + 6\ln(A/\sh) = 21\ \ . \lab{cond1} \end{equation} For consistency, we must require that $A$ not be small, so that for $A\ge 1,$ one obtains an upper bound \be A/\hat\sigma\le 28\ \ . \lab{ash} \end{equation} As before, I now proceed to calculate the number of (hidden) hadrons in a critical bubble. With the same spectrum of glueballs as for the non-SUSY example of the last section (Eq.~\eqn{nc2} with a factor of 4 included for the supermultiplet), and Eq.~\eqn{rcf} for the bubble radius, one finds \be N_c=(\sqrt{6}/\pi^2)\ (T_c/\mu)^2\ f(M_0/T_c)\ (A/\sh)^{3/2} \lab{nch} \end{equation} where again $1/\mu^2$ is the Regge slope for the hidden sector glueballs. As previously, the requirement that the cosmological description be thermodynamically viable requires that $N_c\gg 1,$ which I take to mean $N_c\,\raisebox{-0.13cm}{$\stackrel{\textstyle>}{\textstyle\sim}$}\, 100.$ With the use of \eqn{ash} this devolves to a constraint on $f(M_0/T_c),$ and hence on $M_0:$ for $\mu/T_c\ge 3,$ I find \be M_0\le 1.4\ T_c\ \ , \lab{mbound1} \end{equation} which is of dubious credibility since we expect $M_0>\mu.$ For $\mu/T_c\ge 2,$ the bound is \be M_0\le 2.0\ T_c\ \ , \lab{mbound2} \end{equation} which is marginally possible. Thus, the conclusion at this point is that a bubble description for the first order transition is barely possibly (for $M_0\simeq \mu\simeq 2\ T_c),$ but seemingly unlikely. \section{Hidden Sector with Matter Fields} Suppose that the hidden sector contains $N_f< N$ flavors of vector-like pairs of chiral superfields $Q+\overline{Q}.$ If these are massive ($M_Q\gg \Lambda_N,$ the \su{N} confining scale), then they effectively decouple from the dynamics discussed in this paper. If they are massless (or nearly massless), then the discussion in Refs.~\cite{Ama} and \cite{Aff} is germane: the existence of flat directions $v_{ir}=v_r\delta_{ir} (i=\mbox{color}, r=\mbox{flavor})$ in the field space of the $Q+\overline{Q}$ breaks the symmetry to \su{N^{\, \prime}}, $N^{\, \prime}\equiv N-N_f,$ at the scale $v.$ Between $v$ and $\Lambda_{N^{\, \prime}},$ the effective massless degrees of freedom are the Goldstone bosons of the broken flavor symmetry and the \su{N^{\, \prime}} gauge degrees of freedom. (I assume that it is $\Lambda_{N^{\, \prime}}$ which establishes the condensate scale of interest in this work.) If the Goldstones are in thermal equilibrium with the \su{N^{\, \prime}} fields, then they would contribute to the pressure of the critical bubble with high number density, and the problems encountered earlier would be alleviated. I will now show that the Goldstones are {\em not} in thermal equilibrium with the \su{N^{\, \prime}} fields, and thus the bubble is transparent to their existence. Thermal equilibrium requires that the ratio $\Gamma/H$ be $>1$ during the era of interest, where $\Gamma$ is the reaction rate of the Goldstones in the \su{N^{\, \prime}} plasma. The coupling of a Goldstone to a pair of \su{N^{\, \prime}} gluons is $\frac{g^2}{32\pi^2v\sqrt{N^{\, \prime}}},$ and the cross section in the plasma is easily calculated: \be \sigma\sim \frac{g^2}{4\pi}\ \left(\frac{g^2}{32\pi^2v}\right)^2\ \ , \lab{cross} \end{equation} independent of temperature and $N^{\, \prime}.$ The Hubble constant $H\simeq \sqrt{\cn_{\rm tot}}\ T^2/M_{\rm Pl},$ so that \be \frac{\Gamma}{H}=\frac{\sigma n v_G}{H}\simeq 10^{-8}\ \frac{{\cal N}_{N^{\, \prime}}}{\sqrt{\cn_{\rm tot}}}\ \frac{TM_{\rm Pl}}{v^2}\ \ , \lab{gamh1} \end{equation} where $n=$ plasma number density, $v_G=$ Goldstone velocity. For the \su{6} example $(N^{\, \prime}=6),$ with $T=T_c\simeq 10^{14}, {\cal N}_{N^{\, \prime}}=131.25, \cn_{\rm tot}=345,$ one finds \be \Gamma/H \simeq 0.007\ (v/T_c)^{-2}\ll 1\ \ \lab{gamh2} \end{equation} for any $v\ge T_c.$ Thus, the Goldstones decouple from the \su{N^{\, \prime}} plasma, and do not contribute to the bubble dynamics. \section{Summary and Conclusions} \begin{itemize} \itm{a}The field theoretic description of hidden gluino condensates must imply a parallel thermal/cosmological description of the phase transition beween the unconfined and confined phases of the unbroken Yang-Mills theory. This work has examined the conditions under which a first order transition in terms of classical bubble nucleation is possible. The result found is that only if the mass of the confined phase glueballs (and superpartners) is very near to $ 2 T_c$ , are critical bubbles large enough to contain an adequate number of quanta of the confined phase particles to satisfy the thermodynamic conditions for a first order transition in the expanding universe. This is true whether or not there are Goldstones associated with massless vector-like matter fields. The highly restrictive conditions on glueball mass leads one to question whether a first order transition is possible in the expanding universe. \itm{b}If a first order transition is not feasible, then a field theoretic description of a second (or higher) order transition may be of interest. This entails some difficulty with the Witten index theorem\cite{Wit}: at the critical temperature, the order parameter (presumably the gaugino condensate) must change in a continuous manner from zero to a non-zero value. However, the index for the final state is $N$ (for \su{N}), whereas for the initial state it is (presumably) zero. Resolution of this problem will no doubt involve some non-trivial input to the effective theory. \itm{c}The critical input to the present analysis is the ratio $M_0/T_c,$ where $M_0$ is the mass of the lightest glueball. It would be extremely useful to have some indication, possibly from a lattice study, of this quantity. A continued pursuit of SUSY on the lattice, following the initial effort in Ref.~\cite{Mon} would be very welcome.\ei \clearpage \noindent{\bf Acknowledgement} This work was supported in part by Grant No. PHY-9411546 from the National Science Foundation.
proofpile-arXiv_065-475
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} The influence of disorder on the Abrikosov vortex lattice in the mixed phase of high-temperature superconductors, such as ${\rm Y}{\rm Ba}_{2}{\rm Cu}_{3}{\rm O}_{7-\delta}$ (YBCO), is an issue of immediate technological interest because pinning of the flux lines by disorder opens the possibility of regaining a dissipation-free current flow in the mixed phase. The flux line (FL) array in a high-temperature superconductor (HTSC) is extremely susceptible to thermal and disorder-induced fluctuations due to the interplay of several parameters, namely the high transition temperature $T_{c}$, large magnetic penetration depths $\lambda$ and short coherence lengths $\xi$, and a strong anisotropy of the material. This leads to the existence of a variety of fluctuation dominated phases of the FL array and very rich phase diagrams for the HTSC materials~\cite{Bl}. We want to consider here the pinning of FLs by point defects such as the oxygen vacancies, which is usually referred to as point disorder. It is well-known that the FL lattice is unstable to point disorder~\cite{LO}. It has been conjectured that due to a {\em collective pinning} by the point disorder, the FL array may form a {\em vortex glass phase} with zero linear resistivity~\cite{F,FGLV,Na,FFH}. Although the existence of a vortex glass (VG) phase has been verified experimentally~\cite{Ko89,Ga91,Saf92,Yeh93,Saf93}, its large scale properties characterizing the nature of the VG phase are still under debate~\cite{Bl}. A possible scenario for a description of the low-temperature properties of the FL array subject to point disorder is the existence of a topologically ordered, i.e., {\em dislocation-free} VG phase, the so-called {\em Bragg glass} phase~\cite{GL} as a thermodynamically stable phase. In this glassy phase, a quasi long range positional order of the FL array is maintained in spite of the pinning~\cite{Na,GL}. This entails the existence of algebraically decaying Bragg peaks in diffraction experiments on this phase (bearing the name ``Bragg glass'' for this property), which have indeed been observed in neutron diffraction experiments on ${\rm Bi}_{2}{\rm Sr}_{2}{\rm CaCu}_{2}{\rm O}_{8+\delta}$ (BSCCO) at low magnetic fields~\cite{Cub93}. In the Bragg glass phase, the disordered FL array is modeled as an elastic manifold in a periodic random potential, similar to a randomly pinned charge density wave or a XY model in a random field~\cite{VF,CO,K,GL}. To give a thermodynamically stable phase, this requires the persistence of the topological order, or absence of unbound dislocations, even in the presence of disorder. In the neutron diffraction experiments by Cubitt et al.~\cite{Cub93}, it has been observed that upon increasing the magnetic field, the Bragg peaks vanish, indicating an instability of the Bragg glass phase. Critical current measurements of Khaykovich et al.~\cite{Kha96} show a sharp drop in the (local) critical current $j_c$ upon decreasing the magnetic induction below a critical value. This can be attributed to the ``disentanglement'' of FLs in the absence of dislocations when topological order is regained and dislocation loops vanish upon lowering the magnetic field. Similarly, changes in the I-V characteristics of YBCO based superlattices below a critical magnetic field can be interpreted as stemming from a sharp drop of the pinning energy and indicate a restoration of positional order in the VG~\cite{O}. The existence of a topological transition has also been demonstrated in recent numerical studies~\cite{GH,RKD}. In the closely related 3D XY model in a random field, vortex loops occur in a topological phase transition at a critical strength of the random field~\cite{GH}. In simulations of disordered FL arrays~\cite{RKD}, a proliferation of dislocation lines has been found at a critical magnetic field in good agreement with the experimental results in Ref.~\cite{Cub93}. Recently, the quantitative aspects of this issue have been addressed also analytically~\cite{KNH,F97,EN,Gold,GL97}. In Ref.~\cite{KNH} a self-consistent variational calculation and a scaling argument are presented, which show the topological stability of the elastic Bragg glass phase over a finite range of parameters that can be estimated by a Lindemann-like criterion (\ref{cond}). A more detailed discussion of the methods used in Ref.~\cite{KNH} and their limitations as well as of the Lindemann-like criterion (\ref{cond}), which provides the basis for the calculations in the present work, will be given in the next paragraph. Recently, Fisher~\cite{F97} has presented refined scaling arguments further supporting the existence of a topologically ordered Bragg glass phase. In Refs.~\cite{EN,Gold,GL97}, purely phenomenological Lindemann criteria are used as starting point for an estimate of the phase boundaries of the Bragg glass. Erta\c{s} and Nelson~\cite{EN} and Goldschmidt~\cite{Gold} use ``cage models'' to mimic the interactions between FLs which yields an effective theory for a single FL in a random potential, to which they apply the conventional phenomenological Lindemann criterion. Giamarchi and Le Doussal~\cite{GL97} apply a slightly modified phenomenological Lindemann criterion of the form $\overline{\langle u^2 \rangle}(l) < c_L^2 l^2$, where $\overline{\langle u^2 \rangle}(l)$ is the (disorder-averaged) {\em relative} mean square displacement of neighbouring FLs separated by the FL spacing $l$ ($c_L$ is the Lindemann-number). However, it is well known that phenomenological Lindemann criteria as used in Refs.~\cite{EN,Gold,GL97} do not allow a theoretical description of a phase transition but can only give estimates of the location of the transition. They rely on the assumption that the phase transition reflects in the short scale behaviour of the system. Also the variational calculation and the scaling argument presented in Ref.~\cite{KNH} cannot give a complete description of the transition as only a detailed renormalization group (RG) analysis of the problem would allow which is not yet available. The refined scaling arguments of Ref.~\cite{F97} represent a further step towards this goal. In Ref.~\cite{KNH}, the stability of the elastic Bragg glass phase has been investigated for a layered uniaxial geometry, where the magnetic field is parallel to the CuO-planes, by means of a self-consistent variational calculation and a scaling argument identifying the shear instability due to proliferating dislocations by the disorder-induced decoupling of the layers. For this geometry a Lindemann-like criterion has been derived, which is given below in (\ref{cond}) and relates the stability of the Bragg glass phase to the ratio of the positional correlation length and the FL spacing. These findings are supported by a more rigorous RG analysis \cite{CGL96,KH96} for a simplified model with only two layers of FLs in a parallel magnetic field. The usual experimental situation with the magnetic field perpendicular to the CuO-layers, which will be considered in the present work, is theoretically less understood mainly because displacements of the FLs have two components (biaxial) instead of one component (uniaxial). It is unclear at present whether topological phase transitions of the biaxial and uniaxial model share the same universality class \cite{F97}. In Ref.~\cite{KNH}, it has been argued that the scaling argument for the uniaxial geometry can be generalized to the full, biaxial model leading to the following criterion estimating the range of stability of the Bragg glass phase with respect to a spontaneous formation of dislocation loops: \begin{equation} \label{cond} R_l > c^{1/2\zeta} \left(l^2 +\lambda^2\right)^{1/2} \simeq c^{1/2\zeta} \max{\left\{l,\lambda\right\}}~. \end{equation} $l$ is the FL distance and $\lambda$ the magnetic penetration depth (we consider a magnetic field perpendicular to the CuO-planes of the HTSC, and will specify $\lambda$ below for such a geometry). $R_l$ is the (transversal) {\em positional correlation length} of the disordered FL array, which is defined as the crossover length to the asymptotic large scale behaviour of the Bragg glass phase, where the average FL displacement starts to exceed the FL spacing $l$, see (\ref{Rldef}). $c$ is a number, which was obtained in Ref.~\cite{KNH} to be of the order of $c \approx {\cal O}(50)$, and $\zeta \approx 1/5$ is the roughness exponent of the pre-asymptotic so-called ``random manifold'' regime, see (\ref{zeta}) below. At the boundaries of the regime given by (\ref{cond}), a topological transition occurs, and dislocations proliferate. Beyond the transition line the FL array may form an amorphous VG with vanishing shear modulus or a viscous FL liquid. This article is divided into three parts. First, we will review the pre-asymptotic regimes of the FL array subject to point disorder on scales smaller than the positional correlation length $R_{l}$. This allows us to express $R_l$ in terms of the microscopic parameters of the HTSC and the disorder strength, and to obtain its dependence on magnetic induction $B$ and temperature $T$. In the second part, we will demonstrate the equivalence of the above criterion (\ref{cond}) to the phenomenological Lindemann criterion in the form $\overline{\langle u^2 \rangle}(l) < c_L^2 l^2$ (as for example used in Refs.~\cite{GL97}), where $\overline{\langle u^2 \rangle}(l)$ is the (disorder-averaged) {\em relative} mean square displacement of neighboured FLs. This yields a relation $c \approx c_L^{-2}$ between $c$ from (\ref{cond}) and the Lindemann-number $c_L$, and the value $c \approx {\cal O}(50)$ found by a variational calculation in Ref.~\cite{KNH} turns out to be in good agreement with a value $c_L \approx 0.15$ widely used in the literature for the Lindemann-number. This equivalence further supports a scenario where the topological transition of the FL array subject to point disorder may be described as disorder-induced melting by unbound dislocations on the shortest scale $l$~\cite{KNH}. Finally, and most importantly from the experimental point of view, we estimate the region of the phase diagram of YBCO in the B-T plane (see Fig.~1) where the Bragg glass phase is stable and should be observable experimentally or numerically according to the Lindemann-like criterion (\ref{cond}). We find qualitative agreement with experiments~\cite{O}. The upper phase boundary of the Bragg glass, which we obtain using (\ref{cond}), turns out to be identical to the one obtained by Erta\c{s} and Nelson~\cite{EN}. \section{Positional Correlation Length} To relate the positional correlation length $R_{l}$ to the microscopic parameters of the HTSC and the disorder strength, we have to review the crossover between the different pre-asymptotic regimes of the dislocation-free disordered FL array preceding the asymptotic Bragg glass phase, and the associated crossover length scales~\cite{Bl}. These crossovers are induced by the interplay between the FL interaction, the periodicity of the FL lattice and the disorder potential, which are in addition affected by thermal fluctuations, and lead to essentially two different pre-asymptotic regimes: On the shortest scales, we have the ``Larkin'' or ``random force'' regime of Larkin and Ovchinnikov~\cite{LO}, which crosses over to the so-called ``random manifold'' regime at the Larkin length, before the asymptotic Bragg glass behaviour sets in on the largest scales exceeding the positional correlation length. In between this sequence of crossovers, one additional length scale is set by the FL interaction, which describes a crossover from a ``single vortex'' behaviour to a ``collective'' behaviour. In the following, we consider the usual experimental situation ${\bf H}||{\bf c}$ of a magnetic field perpendicular to the CuO-planes of the HTSC. FL positions are parameterized by the two-component displacement-field ${\bf u}({\bf R},z)={\bf u}({\bf r})$ in a continuum approximation of the Abrikosov lattice, where ${\bf R}$ is the vectors in the ${\bf ab}$-plane and $z$ is the coordinate in the ${\bf c}$-direction, or by the Fourier transform ${\bf \tilde{u}}({\bf K},k_z)={\bf \tilde{u}}({\bf k})$. Let us adopt the convention to denote scales longitudinal to the FLs in the z-direction by $L$ and transversal scales in the ${\bf ab}$-plane by $R$. Moreover, it turns out to be convenient to use the {\em reduced induction} $b \equiv B/B_{c2}(T) = 2\pi \xi_{ab}^2/l^2$ to measure the strength of the magnetic field. \subsection{Interaction-induced Length Scale $L^*$} The dislocation- and disorder-free FL array can be described by elasticity theory (see Ref.~\cite{Bl} for a review) in the displacement field ${\bf u}$ with the elastic moduli $c_{11}$, $c_{44}$ and $c_{66}$, which can in general be dispersive (i.e., k-dependent in Fourier-space) due to the non-locality of the FL interaction. Except for extremely low magnetic fields, the FL lattice is essentially incompressible ($c_{11} \gg c_{66}$), and we can neglect longitudinal compression modes to a good approximation. Note also, that the shear modulus $c_{66}$ is non-dispersive, because volume-preserving shear modes are not affected by the non-locality in the FL interaction. Then, the elastic Hamiltonian in the remaining transversal part ${\bf \tilde{u}}_T$ of the displacement field is of the form: \begin{equation} \label{Hel} {\cal H}_{el}[{\bf \tilde{u}}_T] =\frac{1}{2} \int \frac{d^2{\bf K}}{(2\pi)^2}\frac{dk_z}{2\pi} \left\{ c_{66}({\bf K}\times{\bf \tilde{u}}_T)^2 + c_{44}[K,k_z](k_{z} {\bf \tilde{u}}_T)^2 \right\} \end{equation} The dispersion-free shear modulus is given by $c_{66}\approx \epsilon_0/4l^2$ in the dense limit $l/\lambda_{ab} = (b/2\pi)^{-1/2}/\kappa < 1$ (with $\kappa = \lambda_{ab}/\xi_{ab}$), and exponentially decaying $c_{66} \propto \exp{(-l/\lambda_{ab})} \epsilon_0/l^2$ in the dilute limit $l/\lambda_{ab} >1$. $\epsilon_0 = (\Phi_0/4\pi\lambda_{ab})^2$ is the basic energy (per length) scale of the FL. As estimates for YBCO we use $\xi_{ab}(0)\approx 15\mbox{{\AA}}$, $\epsilon_0(0)\xi_{ab}(0) \approx 1300\mbox{K}$ and $\kappa \approx 100$~\cite{Bl}. The tilt modulus $c_{44}=c_{44}[K,k_z]= c_{44}^{b}[K]+c_{44}^{s}[k_z]$ is dispersive with the bulk-contribution \begin{equation} \label{c44b} c_{44}^{b}[K] \simeq \frac{\epsilon_0}{l^2} \frac{K_{BZ}^2\lambda_{ab}^2}{1+ K^2 \lambda_{c}^2+k_z^2\lambda_{ab}^2} \end{equation} dominating in the dense limit well within the Brillouin zone (BZ) $K < K_{BZ} = 2\sqrt{\pi}/l$ (approximated by a circular BZ) and the {\em single vortex tilt modulus} $c_{44}^{s}= c_{44}^{s}[k_z]$~\cite{GK} \begin{eqnarray} c_{44}^{s}[k_z] &=& c_{44}^{s,J} + c_{44}^{s,em}[k_z] \nonumber\\ &\simeq& \frac{\epsilon_0}{l^2} \left( \varepsilon^2 + \frac{1}{\lambda_{ab}^2k_z^2} \ln{(1+\lambda_{ab}^2k_{z}^2)} \right) \label{c44s} \end{eqnarray} dominating in the dilute limit and at the BZ boundaries $K \simeq K_{BZ}$ or on scales $R \simeq l$. $\varepsilon = \lambda_{ab}/\lambda_{c}$ is the anisotropy ratio of the HTSC and approximately $\varepsilon \approx 1/5$ in YBCO~\cite{Bl}. The single vortex tilt modulus has a strongly dispersive contribution $c_{44}^{s,em}[k_z]$ from the electromagnetic coupling and an essentially dispersion-free contribution $c_{44}^{s,J}$ from the Josephson coupling (where we neglect a logarithmically dispersive factor in $c_{44}^{s,J}$, which is of the order unity for the relevant wavevectors $k_z$ and magnetic inductions $b$). The length scale for the onset of dispersion in the bulk contribution $c_{44}^{b}$ is $\lambda_{c}$ because elements of tilted FLs lying in the ab-plane will start to interact on scales $R < \lambda_c$~\cite{BS}. As length scale for the onset of dispersion, $\lambda_{c}$ occurs as well in the Lindemann criterion (\ref{cond}). The contribution from the electromagnetic coupling to the single vortex tilt modulus gives the local result $c_{44}^{s}[0] \simeq \epsilon_0/l^2$ for $k_z < 1/\lambda_{ab}$, but its strong dispersion $c_{44}^{s,em}[k_z] \propto k_z^{-2}$ for $k_z > 1/\lambda_{ab}$ leads to its suppression at small wavelengths $k_z > 1/\varepsilon\lambda_{ab}$ where $c_{44}^{s} \simeq c_{44}^{s,J}$. From the competition of tilt and shear energy in (\ref{Hel}), we can obtain a scaling relation between scales $L$ longitudinal to the FLs and transversal scales $R$ for typical fluctuations involving elastic deformation: \begin{equation} \label{aspect} L \simeq R \left(\frac{c_{44}[1/R,1/L]}{c_{66}}\right)^{1/2}~. \end{equation} The three-dimensional elastic Hamiltonian (\ref{Hel}) is valid only on scales $R > l$ or \begin{equation} L > L^* \simeq \left( \frac{c_{44}^{s}[1/L^*]}{c_{66}}\right)^{1/2} ~. \label{L*def} \end{equation} When we consider fluctuations on scales $L < L^*$ or $R < l$, the FL array breaks up into single FLs described by 1-dimensional elasticity in the longitudinal coordinate $z$ with a line stiffness $\epsilon_l[k_z] = c_{44}^{s}[k_z]l^2$, because the shear energy containing the effects of FL interactions is always small compared to the tilt energy of the single FL. Thus the interaction-induced length scale $L^*$ separates a regime of ``collective'' behaviour described by 3D elasticity from a ``single vortex'' behaviour described by 1D elasticity. $L^*$ starts to increase exponentially in the dilute limit $l/\lambda_{ab}>1$ due to the exponential decay of $c_{66}$. For the length scale $L^*$ given by (\ref{L*def}), we use therefore the local result $c_{44}^{s} \approx c_{44}^{s,em} \simeq \epsilon_0/l^2$ determined by the electromagnetic coupling. In the dense limit $l/\lambda_{ab}<1$, the scale $L^*$ is smaller than $\varepsilon\lambda_{ab}$, and $c_{44}^{s} \approx c_{44}^{s,J} \simeq \epsilon_0\varepsilon^2/l^2$, i.e., the dispersion-free contribution from the Josephson coupling dominates. This yields \begin{eqnarray} L^* &\approx& \left\{ \begin{array}{ll} l<\lambda_{ab}:&\varepsilon l \\ l>\lambda_{ab}:& l \left( \frac{\lambda_{ab}}{l}\right)^{3/4} \exp{\left(\frac{l}{2\lambda_{ab}}\right)} \end{array} \right. \label{L*} \end{eqnarray} for the interaction induced length scale $L^*$ in the dense and dilute limits. As we will show below, the criterion (\ref{cond}) is indeed equivalent to a Lindemann criterion in a more conventional form where fluctuations $\overline{\langle u^2 \rangle}$ on the transversal scale $R \simeq l$ are considered, see (\ref{Linde}). Therefore, the topological phase transition can be detected by considering fluctuations of single FLs on the longitudinal scale $L \simeq L^*$. We focus in this article on the upper branch of the topological transition line in moderately anisotropic compounds as YBCO such that the electromagnetic coupling can essentially be neglected. However, similar to the findings for thermal melting \cite{BGLN}, the electromagnetic coupling and its strongly dispersive contribution to the single vortex tilt modulus plays an important role for the disorder-induced topological phase transition in very anisotropic compounds such as BSCCO~\cite{remark}. We will consider effects from the electromagnetic coupling in detail in Ref.~\cite{JKF2}. We mention here only that in moderately anisotropic HTSC compounds with the upper branch $b_{t,u}(T)$ of the topological phase transition line will lie entirely in the dense regime $b> 2\pi/\kappa^2$, but below the so-called ``crossover field'' $b_{cr} \sim (2\pi/\kappa^2) (\varepsilon\lambda_{ab}^2/d^2)$ above which $L^*< d$ and the layered structure of the HTSC becomes relevant at the transition line and requires a discrete description in the ${\bf c}$-direction. Only in this regime of magnetic inductions, the strongly dispersive electromagnetic contribution can be neglected at the topological transition (because $L^*<\varepsilon\lambda_{ab}$), while a continuous description in the ${\bf c}$-direction still applies. In the very anisotropic Bi-compounds, however, the upper branch $b_{t,u}(T)$ of the topological phase transition line typically lies in the dilute limit $b< 2\pi/\kappa^2$ where the electromagnetic coupling gives the relevant, strongly dispersive contribution to the single vortex tilt modulus $c_{44}^{s} \approx c_{44}^{s,em}[k_z] \propto k_z^{-2}$. Because of this dispersion, the behaviour of a single vortex of length $L^*$ changes drastically, and short-scale fluctuations on the (longitudinal) scale $L \simeq \max{\{\varepsilon\lambda_{ab},d\}}$ give the main contribution~\cite{JKF2}. In the following, we consider the dense regime of a moderately anisotropic compounds such as YBCO and can thus neglect the dispersive electromagnetic contribution and use the dispersion-free, anisotropic result $c_{44}^{s} \approx c_{44}^{s,J} \simeq \epsilon_0 \varepsilon^2/l^2$. At the lower branch of the topological transition line in the dilute limit (where $L^* > \lambda_{ab}$), we have to take into account the electromagnetic coupling and use the isotropic contribution $c_{44}^{s} \approx c_{44}^{s,em}[0] \simeq \epsilon_0/l^2$ in the local limit. Furthermore, also in this regime effects from the strong dispersion of $c_{44}^{s,em}[k_z]$ have to be considered. The details of the calculation of the lower branch of the topological transition line will be given in Ref.~\cite{JKF2}, and we will mention only the main results below. \subsection{Larkin Length} When point disorder is introduced, every vortex at position ${\bf R_{\nu}}$ in the Abrikosov lattice experiences a pinning potential $V({\bf r})$ with mean zero and short-range correlations \begin{equation} \overline{V({\bf r})V({\bf r}')} = \gamma \xi_{ab}^4 \delta^2_{r_T}({\bf R}- {\bf R}')\delta^{1}_{\xi_{c}}(z-z')~, \end{equation} where the overbar denotes an average over the quenched disorder. The strength of the disorder potential is given by $\gamma= n_{pin}f_{pin}^2$, where $n_{pin}$ is the density of pinning centers and $f_{pin}$ the maximum pinning force exerted by one pinning center, and the effective range of the disorder potential is given by \begin{equation} r_T = \left(\xi_{ab}^2 + \langle u^2 \rangle_{th}(0,L_{\xi})\right)^{1/2} \label{rT} \end{equation} ($\langle\ldots\rangle_{th}$ denotes a purely thermal average for a fixed $V({\bf r})$), which is equal to the size $\xi_{ab}$ of the core of a vortex at $T=0$ but broadened by thermal fluctuations at higher temperatures. As proposed in Ref.~\cite{Bl}, we introduce the dimensionless disorder strength $\delta$ as \begin{equation} \delta ~=~ \frac{\gamma\xi_{ab}^3}{(\epsilon_0\xi_{ab})^2}~. \end{equation} The interaction with the disorder is described by the Hamiltonian \begin{equation} {\cal H}_{dis}[{\bf u}] ~=~ \sum_{\bf \nu} \int dz V({\bf R_{\nu}}+ {\bf u}({\bf R_{\nu}},z),z)~. \label{Hdis} \end{equation} For mean square displacements \begin{equation} \label{conv} u(R,L) \equiv \overline{\langle ({\bf u}({\bf r} + ({\bf R},L))- {\bf u}({\bf r}))^2 \rangle}^{1/2} \end{equation} smaller than the effective scale $r_T$ for variations of the disorder potential $V$, the FLs explore only {\em one} minimum of the disorder potential and perturbation theory in the displacements is valid. Expanding in (\ref{Hdis}) the disorder potential $V$ in ${\bf u}$ yields the {\em random force} theory of Larkin and Ovchinnikov~\cite{LO}. The (longitudinal) {\em Larkin length} $L_{\xi}$ is defined as the crossover scale for the random force regime, at which the average FL displacement becomes of order of the effective range $r_T$ of the point disorder: \begin{equation} \label{Lxidef} u(0,L_{\xi}) \simeq r_T~. \end{equation} It is important to note that for HTSCs such as YBCO and BSCCO, the generic disorder strength is such that \begin{equation} \label{single} L_{\xi} < L^* \end{equation} in the range of magnetic inductions where the elastic Bragg glass will turn out to be stable~\cite{remark2}. Therefore, the random force regime lies entirely in the single vortex regime defined above, and the Larkin length $L_{\xi}$ is given by the single vortex result $L_{\xi}^{s}$, which is at low temperatures~\cite{Bl} \begin{equation} \label{Lxi0} L_{\xi}^{s}(0) \simeq \varepsilon \left( \frac{(\varepsilon\epsilon_0 \xi_{ab})^2} {\varepsilon\gamma} \right)^{1/3} \simeq \varepsilon\xi_{ab}~\left(\frac{\delta}{\varepsilon}\right)^{-1/3}~. \end{equation} This result holds as long as $r_T \simeq \xi_{ab}$. However, above the {\em depinning temperature} $T_{dp}^{s}$ of the single vortex, $r_T$ grows beyond $\xi_{ab}$~\cite{Bl}: \begin{equation} \label{r_T} r_T^2 \simeq \xi_{ab}^2 \left(1 +\exp{\left( \left(T/T_{dp}^{s}\right)^3 \right)}\right)~, \end{equation} where the depinning temperature $T_{dp}^{s}$ is given by~\cite{Bl} \begin{equation} \label{Tdp} T_{dp}^{s} ~\simeq~ \varepsilon\epsilon_0 \xi_{ab}~\frac{\varepsilon\xi_{ab}}{L_{\xi}^{s}(0)} ~\simeq~ \varepsilon\epsilon_0\xi_{ab}~\left(\frac{\delta}{\varepsilon}\right)^{1/3}~. \end{equation} Above $T_{dp}^{s}$, $L_{\xi}^{s}(T)$ increases exponentially with temperature due to the fact that random forces are only marginally relevant for a single FL with two-component displacements~\cite{Bl}: \begin{equation} \label{LxiT} L_{\xi}^{s}(T) \simeq L_{\xi}^{s}(0) \left\{ \begin{array}{ll} T\ll T_{dp}^{s}:~& 1 \\ T>T_{dp}^{s}:~& \left(T/T_{dp}^{s}\right)^{-1} \exp{\left(\left(T/T_{dp}^{s}\right)^3\right)} \end{array} \right. \end{equation} Let us discuss estimates of the quantities $L_{\xi}$ and $T_{dp}^{s}$ at this point, which provide alternative measures of the disorder strength for a HTSC. In Ref.~\cite{EN}, the disorder strength is given by $T_{dp}^{s}\approx 10K$ in BSCCO (where $\varepsilon\approx 1/100$), which leads to $\delta/\varepsilon \approx 1$ with (\ref{Tdp}). This estimate is considerably higher than typical values given in Ref.~\cite{Bl} for weak pinning. Therefore, we will use instead estimates in the range $\delta/\varepsilon \approx 10^{-3}\ldots 10^{-1}$ for YBCO in accordance with Ref.~\cite{Bl} which yield $T_{dp}^{s} \approx 20\ldots 65\mbox{K}$ and values of the order of $L_{\xi}^{s}(0) \approx 30\ldots 6\mbox{{\AA}}$ for the (longitudinal) Larkin length in YBCO. For higher disorder strengths the $T=0$ Larkin length $L_{\xi}^{s}(0)$ can become {\em smaller} than the layer spacing $d$. In YBCO, where $d \approx 12\mbox{{\AA}}$, this happens for quite strong disorder $\delta/\varepsilon \gtrsim 2 \cdot 10^{-2}$. Then, each pancake-like FL element of length $d$ is pinned individually and we enter a {\em strong pinning} regime. This requires a description of pinning at the scale $d$ as the smallest physical length scale in the longitudinal direction. In other words, we have to consider the disorder-induced relative displacement $r_{d}^2 = \overline{\langle u^2 \rangle}$ of two pancake vortices in adjacent layers. This has been calculated in Ref.~\cite{KGL} at $T=0$ by means of an Imry-Ma argument (see also~\cite{engel}) with the result \begin{equation} r_{d}^2(0) \approx \frac{d U_p }{\epsilon_0\varepsilon^2 \ln{\left(d^2/\varepsilon^2 r_{d}^2(0)\right)}} \ln^{-1/2}{\left( \frac{ r_{d}^2(0) }{2\sqrt{\pi}\xi_{ab}^2}\right)}~, \label{rd(0)} \end{equation} where we introduced the mean-square disorder energy $U_p^2 := \gamma d\xi_{ab}^2$ of a line-segment of length $L \simeq d$. The result (\ref{rd(0)}) is valid for $r_d(0) > \xi_{ab}$, i.e., if the relative displacement exceeds the correlation length of the disorder potential, which is the case just for $d> L_{\xi}^{s}(0)$. The equation (\ref{rd(0)}) has to be solved self-consistently, but in the following we will use the estimate obtained in the zeroth iteration \begin{equation} r_{d}^{2}(0) \simeq \frac{d U_p }{\epsilon_0 \varepsilon^2} \simeq \xi_{ab}^2\frac{U_p}{T_{dp}^{s}} \simeq \xi_{ab}^2 \left(\frac{d}{L_{\xi}^{s}(0)}\right)^{3/2}~. \label{rd0(0)} \end{equation} The exponent $3/2= 2\zeta(1,0)$, see below (\ref{rough1}), can be interpreted as the exponent characterizing the end-to-end displacement of a rigid rod that can tilt in a random potential. On scales $L_{\xi}^{s}(0) < L < d$, each pancake can be treated as such a rigid rod of length $L$. Because the pinning is strong, each pancake remains individually pinned in the presence of thermal fluctuations until the thermal energy $T$ is greater than the typical pinning energy $U_p$ of each pancake. Therefore, the result (\ref{rd0(0)}) remains to a good approximation valid in the whole temperature range $T \le U_p$: $r_{d}(T) \simeq r_{d}(0)$. This can be checked in a variational calculation along the lines of Ref.~\cite{engel}. For YBCO with a disorder strength $\delta/\varepsilon \approx 2\cdot 10^{-1}$ we find $U_p \simeq 70\mbox{K}$. Although the thermally increased Larkin length $L_{\xi}^{s}(T)$ becomes equal to the layer spacing $d$ already at a temperature \begin{equation} T_{L_\xi=d} \simeq T_{dp}^{s} \left(1-\frac{L_{\xi}^{s}(0)}{d}\right) < U_p~, \label{TLxid} \end{equation} [in YBCO, we find $T_{L_\xi=d} \simeq 45\mbox{K}$ for $\delta/\varepsilon \approx 2\cdot 10^{-1}$], the crossover from the strong pinning on the scale $d$ to collective pinning on the scale of the Larkin length $L_{\xi}^{s}(T)$ can happen only at $T \simeq U_p$, where the strongly pinned individual pancakes can thermally depin. This result can be obtained from Ref.~\cite{engel}, where it is shown that perturbation theory gives only a {\em locally} stable solution in a variational treatment of the pinning problem for two pancake vortices in adjacent layers in the temperature range $T_{L_\xi=d} < T < U_p$ whereas the result (\ref{rd0(0)}) represents the {\em globally} stable solution. \subsection{Positional Correlation Length $R_l$} On scales exceeding the Larkin length $L_{\xi}^s$, the FLs start to explore many minima of the disorder potential $V$. However, as long as $u(R,L)$ is smaller than the FL spacing $l$, FLs are {\em not} competing for the same minima, and the FLs experience effectively {\em independent} disorder potentials. This leads to the approximation ${\cal H}_{dis}[{\bf u}] \approx \int d^3{\bf r}\tilde{V}({\bf r},{\bf u}({\bf r}))$ (on longitudinal scales exceeding the layer spacing $d$), where $\tilde{V}$ has also short-range correlations in ${\bf u}$. This regime is referred to as the {\em random manifold} regime~\cite{GL}. For a d-dimensional (dispersion-free) elastic manifold with a n-component displacement field ${\bf u}$, the scaling behaviour of the $\overline{\langle uu\rangle}$-correlations is known to be \begin{equation} \label{rough1} u(0,L) \sim L^{\zeta(d,n)} \end{equation} with a roughness exponent $\zeta(d,n)$. We are interested here in the case $d=1$, $n=2$, which is realized on scales $d,L_{\xi}^{s} < L < L^*$ in the single vortex regime, where the FLs are described as 1-dimensional elastic manifolds, and the case $d=3$, $n=2$ on scales $L^* < L <L_{l}$ (or transversal scales $l < R < R_{l}$) in the collective regime, where the FL array is described as 3-dimensional elastic manifold. $L_{l}$ and $R_{l}$ are the {\em positional correlation lengths}, which are defined as the crossover scales for the random manifold regime, at which the average FL displacement becomes of the order of the FL distance $l$: \begin{equation} \label{Rldef} u(R_{l},0) = u(0,L_l) = l~. \end{equation} On scales $R > R_{l}$, where $u(R) > l$, FLs start to compete for the {\em same} minima, and the periodicity of the FL lattice becomes crucial~\cite{Na,GL}. The FL array reaches its asymptotic behaviour of the Bragg glass phase with only logarithmically diverging $\overline{\langle uu\rangle}$-correlations, i.e., quasi long range positional order. The best estimates available for the roughness exponents are~\cite{HZ} \begin{equation} \label{zeta} \zeta(1,2) \approx 5/8 \qquad\mbox{and}\qquad \zeta\equiv \zeta(3,2) \approx 1/5~, \end{equation} where the latter occurs also in the above Lindemann criterion (\ref{cond}). In the collective regime the scaling relation (\ref{rough1}) gets slightly modified by the dispersion (\ref{c44b}) of $c_{44}^{b}$ to \begin{equation} \label{rough2} u^2(R,0) \sim \left(\lambda_{c}^2 + R^2\right)^{\zeta(3,2)}~, \end{equation} as can be checked by means of a simple Flory-type argument, where we equate the typical disorder energy and elastic energy (\ref{Hel}) on {\em one} dominant scale. (As suggested by a more elaborate variational calculation as in Ref.~\cite{GL} we neglect possible small logarithmic corrections of order $\ln{(1/\varepsilon)}$ in (\ref{rough2}).) Note that for $l < R < \lambda_{c}$, the relative displacements (\ref{rough2}) are only marginally growing due to the dispersion of $c_{44}^{b}$. The scaling relations (\ref{rough1},\ref{rough2}) enable us to obtain the relation between the (transversal) positional correlation length $R_l$ and the (longitudinal) Larkin length $L_{\xi}^{s}$, which will allow us to express $R_{l}$ in terms of microscopic parameters, both for weak pinning on the scale $L_{\xi}^{s}$ (for $L_{\xi}^{s}(T)>d$) and for strong pinning of pancakes on the scale $d$ (for $L_{\xi}^{s}(T)<d$). Applying the scaling relation (\ref{rough1}) for the $\overline{\langle uu\rangle}$-correlations to the single vortex random manifold regime on longitudinal scales $L_{\xi}^{s} < L < L^*$, we obtain for the case $L_{\xi}^{s}(0)>d$ of weak pinning \begin{equation} \label{LxiL*} u_* \equiv u(l,0) \simeq u(0,L^*) \simeq r_T~\left(\frac{L^*}{L_{\xi}^s(T)}\right)^{\zeta(1,2)}~. \end{equation} In the same manner we can use (\ref{rough2}) in the collective random manifold regime on transversal scales $l < R < R_{l}$: \begin{equation} \label{lRl} l^2 = u^2(R_{l}) \simeq u_*^2 \left( \frac{\lambda_{c}^2+R_l^2}{\lambda_{c}^2 + l^2} \right)^{\zeta(3,2)} \simeq u_*^2 \left( \frac{R_l}{\lambda_{c}} \right)^{2\zeta(3,2)}~ \end{equation} with $R_l \gg \lambda_c \gg l$. Using (\ref{LxiL*},\ref{lRl}), $R_l$ can be expressed as \begin{equation} \label{RlLxi>d} R_{l}(T) \simeq \lambda_{c} \left(\frac{l}{{r_T}}\right)^{1/\zeta(3,2)} \left(\frac{L_{\xi}^s(T)}{L^*}\right)^{\zeta(1,2)/\zeta(3,2)}. \end{equation} With the results (\ref{r_T}) for $r_T$, (\ref{L*}) for $L^*$, and (\ref{LxiT}) for $L_{\xi}^s(T)$ together with $\zeta(3,2) \approx 1/5$ and $\zeta(1,2) \approx 5/8$ (\ref{zeta}), this yields the desired expression for $R_l$: \begin{eqnarray} R_{l}(0) &\approx& \lambda_{c} \left(\frac{b}{2\pi}\right)^{-{15}/{16}} \left(\frac{\delta}{\varepsilon}\right)^{-{25}/{24}} \nonumber \\ R_{l}(T) &\approx& R_{l}(0) \left\{ \begin{array}{ll} T\ll T_{dp}^{s}:& 1 \\ T>T_{dp}^{s}:& \left(T/T_{dp}^{s}\right)^{-{25}/{8}} \exp{\left( \frac{5}{8} \left(T/T_{dp}^{s}\right)^3\right)} \end{array} \right. ~. \label{Rl2Lxi>d} \end{eqnarray} The weakening of the pinning by thermal fluctuations leads to an exponential increase of $R_l(T)$ for temperatures above the depinning temperature $T_{dp}^{s}$ similar (and related) to the behaviour of the thermally increased Larkin length $L_{\xi}^{s}(T)$. For inductions $b/2\pi = 10^{-4}\ldots 10^{-2}$ in the dense limit $b/2\pi > 1/\kappa^2$, a disorder strength $\delta/\varepsilon \approx 10^{-2}$, and $\lambda_c(0) \approx 7500\mbox{{\AA}}$, we obtain (transversal) positional correlation lengths $R_{l}(0) \approx (10^4\ldots 10^6)\cdot\lambda_c \approx 7,5\cdot(10^{-1}\ldots 10)\mbox{cm}$, which are extremely large indicating that over a wide range of length scales the pre-asymptotic random manifold regimes should be observable rather than the asymptotic Bragg glass regime. For the case $L_{\xi}^{s}(0)<d$ of strong pinning of pancakes on the scale $d$, we apply (\ref{rough1}) for the $\overline{\langle uu\rangle}$-correlations to the single vortex random manifold regime on longitudinal scales $d < L < L^*$ and obtain for low temperatures \begin{equation} \label{dL*} u_* \equiv u(R=l,0) \simeq u(0,L=L^*) \simeq r_d(0)~\left(\frac{L^*}{d}\right)^{\zeta(1,2)}~ \end{equation} instead of (\ref{LxiL*}). Using this and (\ref{lRl}), $R_l$ can be expressed as \begin{equation} \label{RlLxi<d} R_{l}(0) \simeq \lambda_{c} \left(\frac{l}{{r_d(0)}}\right)^{1/\zeta(3,2)} \left(\frac{d}{L^*}\right)^{\zeta(1,2)/\zeta(3,2)}. \end{equation} With the results (\ref{rd0(0)}) for $r_d(0)$, (\ref{L*}) for $L^*$, and $\zeta(3,2) \approx 1/5$ and $\zeta(1,2) \approx 5/8$ (\ref{zeta}), we obtain for the positional correlation length $R_l$: \begin{eqnarray} R_{l}(0) &\approx& \lambda_{c} \left(\frac{b}{2\pi}\right)^{-15/16} \left(\frac{\delta}{\varepsilon}\right)^{-5/6} \left(\frac{\xi_{ab}}{d} \varepsilon\right)^{5/8} \label{Rl2Lxi<d} \end{eqnarray} This result is valid for temperatures $T \le U_p$ and gives a temperature independent, smaller value than (\ref{Rl2Lxi>d}) in this temperature range. At $T \simeq U_p$, pancakes can thermally depin for strong pinning, and we expect a crossover to the weak pinning result (\ref{Rl2Lxi>d}) with a pronounced increase of $R_l(T)$ with temperature. In YBCO, strong pinning is realized for $\delta/\varepsilon \gtrsim 2 \cdot 10^{-2}$. For $\delta/\varepsilon \approx 2\cdot 10^{-1}$ and with the layer spacing $d \approx 12\mbox{{\AA}}$, we find $R_{l}(0) \approx 5\cdot(10^2\ldots 10^4)\cdot\lambda_c \approx 3,8\cdot(10^{-2}\ldots 1)\mbox{cm}$ in the induction range $b/2\pi = 10^{-4}\ldots 10^{-2}$ in the dense limit. \section{Lindemann Criterion} Let us now show the equivalence of the Lindemann-like criterion (\ref{cond}) obtained in Ref.~\cite{KNH} to the conventional form of the Lindemann criterion generalized to a disordered system. The Lindemann criterion has been proven as a very efficient phenomenological tool to obtain the thermal melting curves of lattices, e.g.\ the disorder-free FL lattice. There, it is formulated in its conventional form \begin{equation} \label{Lindemann} \langle u^2 \rangle_{th} = c_L^2 l^2~, \end{equation} with a Lindemann-number $c_L \approx 0.1\ldots 0.2$. For the thermal melting of the FL array, the main contributions to the left hand side of (\ref{Lindemann}) come from fluctuations on the {\em shortest} scale, which is in the transverse direction the FL spacing $l$, i.e., $\langle u^2 \rangle_{th} \approx \langle u^2 \rangle_{th}(l,0)$ (note that we apply again a convention like (\ref{conv})). Therefore, the straightforward generalization of (\ref{Lindemann}) to the disorder-induced melting by dislocations is \begin{equation} \label{Linde} \overline{\langle u^2 \rangle}(l,0) \simeq \overline{\langle u^2 \rangle}(0,L^*) \equiv u_*^2 = c_L^2 l^2~, \end{equation} where we consider again fluctuations on the shortest scale $R \simeq l$. In Ref.~\cite{KNH}, one derivation of the criterion (\ref{cond}) was based on a variational calculation for a layered superconductor in a parallel field. There it was found, that unbound dislocations proliferate indeed on the shortest scale at the topological transition described by (\ref{cond}), i.e., in between every layer and thus with a distance $l$. This suggests that the use of the conventional phenomenological Lindemann criterion in the form (\ref{Linde}) might be one possibility to obtain the topological transition line. This can be further justified by showing that the criterion (\ref{cond}), obtained in Ref.~\cite{KNH} on the basis of a scaling argument and a variational calculation for a uniaxial model, is actually {\em equivalent} to the phenomenological Lindemann criterion (\ref{Linde}): Considering the relation (\ref{lRl}) between $u_*$ and $l$, it becomes clear that (\ref{cond}) is the analog of the Lindemann criterion (\ref{Linde}) formulated in terms of the underlying transversal scales rather than the corresponding displacements. Using (\ref{lRl}), the criterion (\ref{cond}) for the stability of the Bragg glass can be written as \begin{equation} \label{Linde2} u_*^2 < c^{-1} l^2~. \end{equation} This is just the above phenomenological Lindemann criterion (\ref{Linde}), and we can identify \begin{equation} \label{ccL} c \approx c_L^{-2}~. \end{equation} We see that the equivalence of the criterion (\ref{cond}) to the phenomenological Lindemann criterion (\ref{Linde}) includes the agreement of the appearing numerical factors: The value for the Lindemann-number $c_L \approx 0.15$, widely used in the literature, produces a good agreement in (\ref{ccL}) with the value $c \approx {\cal O}(50)$ obtained by the variational calculation. This equivalence to a scenario where disorder-induced fluctuations on the shortest scale ``melt'' the FL array favours a first order transition scenario for the topological transition, which could not be excluded in the experiments~\cite{Kha96}. As we will see, the quantity $u_*^2$ is equivalent to the mean square displacement of the ``effective'' FL studied in the ``cage model'' of Erta\c{s} and Nelson~\cite{EN}. They apply the Lindemann criterion directly in its phenomenological form (\ref{Linde}) to the ``caged'' FL. Using (\ref{ccL},\ref{zeta}), we can cast the Lindemann-like criterion (\ref{cond}) into the form \begin{equation} \label{cond2} R_l > c_L^{-1/\zeta} \left(l^2 +\lambda_{c}^2\right)^{1/2}~\approx~ c_L^{-5} \max{\left\{l,\lambda_{c}\right\}}~. \end{equation} \section{Phase Diagram} Let us now address the issue of phase boundaries of the topologically ordered Bragg glass in the B-T plane as they follow from the Lindemann-like criterion (\ref{cond}) in the above form (\ref{cond2}). The results are summarized in Fig.~1. The boundary of the regime given by (\ref{cond2}) defines a {\em topological transition line} $B_{t}(T)$, where dislocations proliferate and the topological order of the Bragg glass phase is lost. The upper branch $b_{t,u}(T)$ of this line can be obtained by applying the expressions (\ref{Rl2Lxi>d}) or (\ref{Rl2Lxi<d}) for the positional correlation length $R_{l}$ in the dense limit $b > 2\pi/\kappa^2$ to the criterion (\ref{cond2}). For weak pinning or $L_{\xi}^{s}(0)>d$ such that we have collective pinning on the scale $L_{\xi}^{s}$, this yields a condition $b < b_{t,u}(T)$ in the b-T plane with \begin{eqnarray} b_{t,u}(0) &\approx& 2\pi \left(\frac{\delta}{\varepsilon}\right)^{-10/9}c_L^{16/3} ~\approx~ 2\pi \left(\frac{\varepsilon\epsilon_0\xi_{ab}}{T_{dp}^{s}}\right)^{10/3} c_L^{16/3} \nonumber\\ b_{t,u}(T) &\approx& b_{t,u}(0) \left\{ \begin{array}{ll} T\ll T_{dp}^{s}:& 1 \\ T>T_{dp}^{s}:& \left(T/T_{dp}^{s}\right)^{-{10}/{3}} \exp{\left(\frac{2}{3} \left(T/T_{dp}^{s}\right)^3\right)}. \end{array} \right. \nonumber\\ \label{bupTLxi>d} \end{eqnarray} Note that the transition line (\ref{bupTLxi>d}) is identical to the one obtained by Erta\c{s} and Nelson~\cite{EN} by applying the conventional phenomenological Lindemann criterion to a ``cage model'' for a single FL (this demonstrates the equivalence of the displacement $u_*$ as defined in (\ref{Linde}) to the average displacement of the ``caged'' FL). Estimates of $b_{t,u}(0)$ strongly depend on the chosen value for the Lindemann number $c_L \approx 0.1\ldots 0.2$. A value $c_L \approx 0.15$ and a disorder strength $\delta/\varepsilon \approx 10^{-2}$ lead to $b_{t,u}(0) \approx 4\cdot 10^{-2}$ or $B_{t,u}(0) \approx 6\mbox{T}$ with $B_{c2}(0) \approx 150\mbox{T}$. For temperatures $T< T_{dp}^{s}$ the transition line is essentially temperature-independent because the mechanism for the proliferation of dislocation loops is purely disorder-driven at low temperatures~\cite{KNH}. For $T > T_{dp}^{s}$ it increases exponentially due to the very effective weakening of the pinning effects by thermal fluctuations in the single vortex regime. For $L_{\xi}^{s}(0)<d$, i.e., strong pinning of pancakes on the scale $d$ (realized for $\delta/\varepsilon \gtrsim 2 \cdot 10^{-2}$ in YBCO), we obtain instead \begin{eqnarray} b_{t,u}(0) &\approx& 2\pi c_L^{16/3} \left(\frac{\delta}{\varepsilon} \right)^{-4/3} \left(\frac{\xi_{ab}}{d} \varepsilon\right)^{2/3} \label{bup0Lxi<d} \end{eqnarray} at low temperatures. This result gives a lower induction for the topological transition than (\ref{bupTLxi>d}) and remains valid up to the temperature $T \simeq U_p$, where pancakes can thermally depin. For $T > U_p$ we expect a pronounced increase of $b_{t,u}(T)$ with temperature and a crossover to the weak pinning result (\ref{bupTLxi>d}), cf.\ Fig.~1. Estimates for $b_{t,u}(0)$ are again very susceptible to changes in the chosen value for the Lindemann number $c_L$. For $c_L \approx 0.15$ and $\delta/\varepsilon \approx 2\cdot 10^{-1}$, we find $b_{t,u}(0) \approx 8\cdot 10^{-4}$ or $B_{t,u}(0) \approx 0,12\mbox{T}$ for YBCO. The estimates for $B_{t,u}(0)$ obtained from (\ref{bupTLxi>d}) and (\ref{bup0Lxi<d}) are in qualitative agreement with the experimental results for the magnetic field where in YBCO based superlattices a change in the I-V characteristics has been observed~\cite{O}. Both (\ref{bupTLxi>d}) and (\ref{bup0Lxi<d}) show that the magnetic induction $b_{t,u}$ at the topological transition decreases for stronger anisotropy or effectively larger disorder strength $\delta/\varepsilon$, and the stability region of the topologically ordered Bragg glass shrinks. At some temperature $T_{x,u}$ above $T_{dp}^{s}$ (and $T_{L_\xi=d}$), the topological transition line $b_{t,u}(T)$ will terminate in the upper branch of the melting curve $b_{m,u}(T)$, which is \begin{equation} b_{m,u}(T) ~\approx~ 2\pi c_L^4 \left(\frac{\varepsilon \epsilon_0 \xi_{ab}}{T}\right)^2 \label{bmuT} \end{equation} in this regime of inductions for the moderately anisotropic compound YBCO~\cite{BGLN}. The temperature $T_{x,u}$ can be determined from the condition that the thermally increased Larkin length $L_{\xi}^{s}(T)$ becomes equal to the scale $L^*$ of the dominant fluctuations at the melting line and the topological transition line. Because $\overline{\langle u^2 \rangle}(0,L_{\xi}^{s}(T)) = \langle u^2 \rangle_{th}(0,L_{\xi}^s(T))$ at the thermally increased Larkin length, the Lindemann criteria (\ref{Lindemann}) for thermal melting and (\ref{cond2}) in the form (\ref{Linde}) for the topological phase transition are indeed fulfilled {\em simultaneously} if $L^* = L_{\xi}^{s}(T)$: \begin{equation} \overline{\langle u^2 \rangle}(0,L^*) = \langle u^2 \rangle_{th}(0,L^*) = c_L^2 l^2~. \end{equation} This yields \begin{equation} T_{x,u} \simeq T_{dp}^{s} \ln^{1/3}{\left( \left(\frac{\delta}{\varepsilon}\right)^{2/3}c_L^{-2}\right)} \label{Txu} \end{equation} In YBCO, we find $T_{x,u} \approx 80\mbox{K}$ with the above estimates $c_L \approx 0.15$ and $\delta/\varepsilon \approx 2\cdot 10^{-1}$. For $T>T_{x,u}$ beyond the melting curve $b_{m,u}(T)$, the FL array melts into a disordered FL liquid, and the Bragg glass order is destroyed by the thermal fluctuations on small scales where disorder-induced fluctuations are irrelevant, whereas above the transition line $b_{t,u}(T)$ the Bragg glass ``melts'' by disorder-induced fluctuations, when unbound dislocation loops proliferate. For $T<T_{x,u}$, we find $b_{m,u}(T) > b_{t,u}(T)$, and the melting curve $b_{m,u}(T)$ lies {\em above} the topological transition line in the b-T plane. Therefore, we expect for temperatures $T<T_{x,u}$ that an amorphous, i.e., topologically disordered vortex glass melts into a vortex liquid at the thermal melting line $b_{m,u}(T)$ and consequently, that the melting transition into a vortex liquid at $b_{m,u}(T)$ is of a different nature below and above $T_{x,u}$. In the experiments reported in Ref.~\cite{Saf93}, such a change in the properties of the melting transition has indeed been observed in YBCO at a temperature around $75\mbox{K}$ which is in fairly good agreement with our result for $T_{x,u}$. Let us now give the main results for the lower branch of the topological transition line $b_{t,l}(T)$ at which the strongly dispersive contribution from the electromagnetic coupling to the single vortex tilt modulus is dominating. At low inductions in the dilute limit $b \ll 2\pi/\kappa^2$, the criterion (\ref{cond2}) will be violated due to the exponential decrease of the shear modulus $c_{66}$, or increase of the interaction-induced length scale $L^*$ (\ref{L*}). At low temperatures $T\simeq 0$, the positional correlation length $R_l(0)$ can be determined also from (\ref{RlLxi>d}) using the {\em isotropic} single vortex Larkin length (given by (\ref{Lxi0}) with $\varepsilon=1$) and the appropriate result for $L^*$. Because the isotropic Larkin length is always greater than the layer spacing, the layered structure is irrelevant for the collective pinning at low inductions. The criterion (\ref{cond2}) then yields for the lower branch of the topological transition line \begin{equation} b_{t,l}(0)\approx \frac{2\pi}{\kappa^2} \ln^{-2}{\left(c_L^{16/5} \kappa^{6/5} \delta(0)^{-2/3}\right)} \label{blow0} \end{equation} Due to the strong dispersion of $c_{44}^{s,em}[k_z]$, the thermal depinning at higher temperatures is more complex and involves several crossover temperatures. However, only above the {\em isotropic} single vortex depinning temperature $T_{dp,i}^{s}$ (given by (\ref{Tdp}) with $\varepsilon=1$) the positional correlation length is increasing exponentially similarly to the thermally increased isotropic Larkin length. This gives only a weak logarithmic temperature dependence for $T<T_{dp,i}^{s}$ whereas we find the asymptotics \begin{equation} \label{blowT} b_{t,l}(T) ~\sim~ \frac{25\pi}{2\kappa^2} \left(\frac{T}{T_{dp,i}^{s}}\right)^{-6}~. \end{equation} at temperatures $T \gg T_{dp,i}^{s}$ well above the isotropic single vortex depinning temperature. At a temperature $T_{x,l}~(>T_{dp,i})$, $b_{t,l}(T)$ will terminate in the lower branch of the melting curve $b_{m,l}(T)$, which increases logarithmically with temperature~\cite{BGLN} \begin{equation} b_{m,l}(T) ~\approx~ \frac{2\pi}{\kappa^2} \ln^{-2}{\left( \frac{ {c_L}^4\epsilon_0^2 \lambda_{ab}^2}{T^2} \right)}~. \label{bmlT} \end{equation} Analogously to the findings for the upper branch of the topological transition line, $T_{x,l}$ can be determined from the condition that the thermally increased isotropic Larkin length becomes equal to the scale $L^*$ at the melting line. This yields \begin{equation} T_{x,l} \simeq T_{dp,i}^{s} \ln^{1/3}{\left(c_L^2 \kappa \varepsilon^{2/3} \left(\frac{\delta}{\varepsilon}\right)^{-1/3} \right)} \label{Txl} \end{equation} With $c_L \approx 0.15$ and $\delta/\varepsilon \approx 10^{-2}$, we obtain $b_{t,l}(0) \approx 0.16 (2\pi/\kappa^2) \approx 1\cdot 10^{-4}$, which is by a factor of 40 smaller than $b_{t,u}(0)$ and experimentally hard to verify due to the small inductions $B_{t,u}(0) \approx 150\mbox{G}$. Furthermore we find $T_{dp,i} \approx 70\mbox{K}$ and $T_{x,l}\approx 85\mbox{K}$. From (\ref{blow0}) it is clear that the transition line $b_{t,u}(T)$ increases with the disorder strength so that the stability region of the topologically ordered Bragg glass shrinks. \section{Conclusion} In conclusion, we have obtained the region in the phase diagram of YBCO in the B-T plane, where the topologically ordered vortex glass should be observable, and the topological transition lines $B_{t,u}(T)$ and $B_{t,l}(T)$, where dislocation loops proliferate. The resulting phase diagram, as given by the formulae (\ref{bupTLxi>d}), (\ref{bup0Lxi<d}), (\ref{blow0}), and (\ref{blowT}) is depicted in Fig.~1. The results are in qualitative agreement with the experimental data of Ref.~\cite{O} if the observed changes in the I-V characteristics are attributed to a topological transition of the disordered vortex array. The phase diagram is based on the Lindemann-like criterion (\ref{cond}) or (\ref{cond2}), which has been obtained by a variational calculation for a uniaxial model and a scaling argument presented in Ref.~\cite{KNH}. We have demonstrated the equivalence to the conventional phenomenological formulation of the Lindemann criterion (\ref{Linde}) up to the involved numerical factors, i.e., the Lindemann-number $c_L$. Our results for the upper branch of the topological transition line $B_{t,u}(T)$ agree with Ref.~\cite{EN}, where the conventional phenomenological Lindemann-criterion was applied to the disorder-induced ``melting'' in the framework of a ``cage model''. \section{Acknowledgments} The author thanks T.~Nattermann, T.~Hwa, and A.E.~Koshelev for discussions and support by the Deutsche Forschungsgemeinschaft through SFB~341~(B8) and grant \mbox{KI~662/1--1}.
proofpile-arXiv_065-476
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} A common theme in higher-dimensional algebra is `categorification': the formation of $(n+1)$-categorical analogs of $n$-categorical algebraic structures. This amounts to replacing equations between $n$-morphisms by specified $(n+1)$-isomorphisms, in accord with the philosophy that any {\it interesting} equation --- as opposed to one of the form $x = x$ --- is better understood as an isomorphism, or more generally an equivalence. In their work on categorification in topological quantum field theory, Freed \cite{Freed} and Crane \cite{Crane} have, in an informal way, used the concept of a `2-Hilbert space': a category with structures and properties analogous to those of a Hilbert space. Our goal here is to define 2-Hilbert spaces precisely and begin to study them. We concentrate on the finite-dimensional case, as the infinite-dimensional case introduces extra issues that we are not yet ready to handle. We must start by categorifying the various ingredients in the definition of Hilbert space. These are: 1) the zero element, 2) addition, 3) subtraction, 4) scalar multiplication, and 5) the inner product. The first four have well-known categorical analogs. 1) The analog of the zero vector is a `zero object'. A zero object in a category is an object that is both initial and terminal. That is, there is exactly one morphism from it to any object, and exactly one morphism to it from any object. Consider for example the category ${\rm Hilb}$ having finite-dimensional Hilbert spaces as objects, and linear maps between them as morphisms. In ${\rm Hilb}$, any zero-dimensional Hilbert space is a zero object. 2) The analog of adding two vectors is forming the direct sum, or more precisely the `coproduct', of two objects. A coproduct of the objects $x$ and $y$ is an object $x \oplus y$, equipped with morphisms from $x$ and $y$ to it, that is universal with respect to this property. In ${\rm Hilb}$, for example, any Hilbert space equipped with an isomorphism to the direct sum of $x$ and $y$ is a coproduct of $x$ and $y$. 3) The analog of subtracting vectors is forming the `cokernel' of a morphism $f \colon x \to y$. This makes sense only in a category with a zero object. A cokernel of $f \colon x \to y$ is an object ${\rm cok} f$ equipped with an epimorphism $g \colon y \to {\rm cok} f$ for which the composite of $f$ and $g$ factors through the zero object, that is universal with respect to this property. Note that while we can simply subtract a number $x$ from a number $y$, to form a cokernel we need to say how the object $x$ is mapped to the object $y$. In ${\rm Hilb}$, for example, any space equipped with an isomorphism to the orthogonal complement of ${\rm im} f$ in $y$ is a cokernel of $f \colon x \to y$. If $f$ is an inclusion, so that $x$ is a subspace of $y$, its cokernel is sometimes written as the `direct difference' $y \ominus x$ to emphasize the analogy with subtraction. An important difference between zero, addition and subtraction and their categorical analogs is that these operations represent extra {\it structure} on a set, while having a zero object, binary coproducts or cokernels is merely a {\it property} of a category. Thus these concepts are in some sense more intrinsic to categories than to sets. On the other hand, one pays a price for this: while the zero element, sums, and differences are unique in a Hilbert space, the zero object, coproducts, and cokernels are typically unique only up to canonical isomorphism. 4) The analog of multiplying a vector by a complex number is tensoring an object by a Hilbert space. Besides its additive properties (zero object, binary coproducts, and cokernels), ${\rm Hilb}$ also has a compatible multiplicative structure, that is, tensor products and a unit object for the tensor product. In other words, ${\rm Hilb}$ is a `ring category', as defined by Laplaza and Kelly \cite{Kelly2,Laplaza}. We expect it to play a role in 2-Hilbert space theory analogous to the role played by the ring ${\Bbb C}$ of complex numbers in Hilbert space theory. Thus we expect 2-Hilbert spaces to be `module categories' over ${\rm Hilb}$, as defined by Kapranov and Voevodsky \cite{KV}. An important part of our philosophy here is that ${\Bbb C}$ is the primordial Hilbert space: the simplest one, upon which the rest are modelled. By analogy, we expect ${\rm Hilb}$ to be the primordial 2-Hilbert space. This is part of a general pattern pervading higher-dimensional algebra; for example, there is a sense in which $n{\rm Cat}$ is the primordial $(n+1)$-category. The real significance of this pattern remains somewhat mysterious. 5) Finally, what is the categorification of the inner product in a Hilbert space? It appears to be the `$\hom$ functor'. The inner product in a Hilbert space $x$ is a bilinear map \[ \langle\,\cdot\,,\, \cdot\, \rangle \colon \overline x \times x \to {\Bbb C} \] taking each pair of elements $v,w \in x$ to the inner product $\langle v,w\rangle$. Here $\overline x$ denotes the conjugate of the Hilbert space $x$. Similarly, the $\hom$ functor in a category $C$ is a bifunctor \[ \hom(\,\cdot\, , \,\cdot\,)\colon C^{\rm op} \times C \to {\rm Set} \] taking each pair of objects $c,d \in C$ to the set $\hom(c,d)$ of morphisms from $c$ to $d$. This analogy clarifies the relation between category theory and quantum theory that is so important in topological quantum field theory. In quantum theory the inner product $\langle v,w \rangle$ is a {\it number} representing the amplitude to pass from $v$ to $w$, while in category theory $\hom(c,d)$ is a {\it set} of morphisms passing from $c$ to $d$. To understand this analogy better, note that any morphism $f \colon x \to y$ in ${\rm Hilb}$ can be turned around or `dualized' to obtain a morphism $f^\ast \colon y \to x$. The morphism $f^\ast$ is called the adjoint of $f$, and satisfies \[ \langle fv,w \rangle = \langle v,f^\ast w \rangle \] for all $v \in x$, $w \in y$. The ability to dualize morphisms in this way is crucial to quantum theory. For example, observables are represented by self-adjoint morphisms, while symmetries are represented by unitary morphisms, whose adjoint equals their inverse. The ability to dualize morphisms in ${\rm Hilb}$ makes this category very different from the category ${\rm Set}$, in which the only morphisms $f \colon x \to y$ admitting any natural sort of `dual' are the invertible ones. There is, however, duals for certain noninvertible morphisms in ${\rm Cat}$ --- namely, adjoint functors. The functor $F^\ast \colon D \to C$ is said to be a right adjoint of the functor $F \colon C \to D$ if there is a natural isomorphism \[ \hom(Fc,d) \cong \hom(c,F^\ast d) \] for all $c \in C$, $d \in D$. The analogy to adjoints of operators between Hilbert spaces is clear. Our main point here is that that this analogy relies on the more fundamental analogy between the inner product and the $\hom$ functor. One twist in the analogy between the inner product and the $\hom$ functor is that the inner product for a Hilbert space takes values in ${\Bbb C}$. Since we are treating ${\rm Hilb}$ as the categorification of ${\Bbb C}$, the $\hom$-functor for a 2-Hilbert space should take values in ${\rm Hilb}$ rather than ${\rm Set}$. In technical terms \cite{Kelly}, this suggests that a 2-Hilbert space should be enriched over ${\rm Hilb}$. To summarize, we expect that a 2-Hilbert space should be some sort of category with 1) a zero object, 2) binary coproducts, and 3) cokernels, which is 4) a ${\rm Hilb}$-module and 5) enriched over ${\rm Hilb}$. However, we also need a categorical analog for the equation \[ \langle v,w\rangle = \overline{\langle w,v \rangle} \] satisfied by the inner product in a Hilbert space. That is, for any two objects $x,y$ in a 2-Hilbert space there should be a natural isomorphism \[ \hom(x,y) \cong \overline{\hom(y,x)} \] where $\overline{\hom(y,x)}$ is the complex conjugate of the Hilbert space $\hom(y,x)$. (The fact that objects in ${\rm Hilb}$ have complex conjugates is a categorification of the fact that elements of ${\Bbb C}$ have complex conjugates.) This natural isomorphism should also satisfy some coherence laws, which we describe in Section 2. We put these ingredients together and give a precise definition of 2-Hilbert spaces in Section 3. Why bother categorifying the notion of Hilbert space? As already noted, one motivation comes from the study of topological quantum field theories, or TQFTs. In the introduction to this series of papers \cite{BD}, we proposed that $n$-dimensional unitary extended TQFTs should be treated as $n$-functors from a certain $n$-category $n{\rm Cob}$ to a certain $n$-category $n{\rm Hilb}$. Roughly speaking, the $n$-category $n{\rm Cob}$ should have 0-dimensional manifolds as objects, 1-dimensional cobordisms between these as morphisms, 2-dimensional cobordisms between these as 2-morphisms, and so on up to dimension $n$. The $n$-category $n{\rm Hilb}$, on the other hand, should have `$n$-Hilbert spaces' as objects, these being $(n-1)$-categories with structures and properties analogous to those of Hilbert spaces. (Note that an ordinary Hilbert space is a `1-Hilbert space', and is a 0-category, or set, with extra structures and properties.) An eventual goal of this series is to develop the framework needed to make these ideas precise. This will require work both on $n$-categories in general --- especially `weak' $n$-categories, which are poorly understood for $n > 3$ --- and also on the particular $n$-categories $n{\rm Cob}$ and $n{\rm Hilb}$. One of the guiding lights of weak $n$-category theory is the chart shown in Figure 1. This describes `$k$-tuply monoidal $n$-categories' --- that is, $(n+k)$-categories with only one $j$-morphism for $j < k$. The entries only correspond to theorems for $n+k \le 3$, but there is evidence that the pattern continues for arbitrarily large values of $n,k$. Note in particular how as we descend each column, the $n$-categories first acquire a `monoidal' or tensor product structure, which then becomes increasingly `commutative' in character with increasing $k$, stabilizing at $k = n+2$. \vskip 0.5em \begin{center} {\small \begin{tabular}{|c|c|c|c|} \hline & $n = 0$ & $n = 1$ & $n = 2$ \\ \hline $k = 0$ & sets & categories & 2-categories \\ \hline $k = 1$ & monoids & monoidal & monoidal \\ & & categories & 2-categories \\ \hline $k = 2$ &commutative& braided & braided \\ & monoids & monoidal & monoidal \\ & & categories & 2-categories \\ \hline $k = 3$ &` & symmetric & weakly involutory \\ & & monoidal & monoidal \\ & & categories & 2-categories \\ \hline $k = 4$ &`' & `' &strongly involutory\\ & & & monoidal \\ & & & 2-categories \\ \hline $k = 5$ &`' &`' & `' \\ & & & \\ & & & \\ \hline \end{tabular}} \vskip 1em 1. The category-theoretic hierarchy: expected results \end{center} \vskip 0.5em At least in the low-dimensional cases examined so far, the $n$-categories of interest in topological quantum field theory have simple algebraic descriptions. For example, knot theorists are familiar with the category of framed oriented 1-dimensional cobordisms embedded in $[0,1]^3$. We would call these `1-tangles in 3 dimensions'. They form not merely a category, but a braided monoidal category. In fact, they form the `free braided monoidal category with duals on one object', the object corresponding to the positively oriented point. More generally, we expect that $n$-tangles in $n+k$ dimensions form the `free $k$-tuply monoidal $n$-category with duals on one object', $C_{n,k}$. By its freeness, we should be able to obtain a representation of $C_{n,k}$ in any $k$-tuply monoidal $n$-category with duals by specifying a particular object therein. When the codimension $k$ enters the stable range $k \ge n + 2$ we hope to obtain the `free stable $n$-category with duals on one object', $C_{n,\infty}$. A unitary extended TQFT should be a representation of this in $n{\rm Hilb}$. If as expected $n{\rm Hilb}$ is a stable $n$-category with duals, to specify a unitary extended TQFT would then simply be to specify a particular $n$-Hilbert space. More generally, we expect an entire hierarchy of $k$-tuply monoidal $n$-Hilbert spaces in analogy to the category-theoretic hierarchy, as shown in Figure 2. We also hope that an object in a $k$-tuply monoidal $n$-Hilbert space $H$ will determine a representation of $C_{n,k}$ in $H$, and thus an invariant of $n$-tangles in $(n+k)$ dimensions. \vskip 0.5em \begin{center} {\small \begin{tabular}{|c|c|c|c|} \hline & $n = 1$ & $n = 2$ & $n = 3$ \\ \hline $k = 0$ & Hilbert & 2-Hilbert & 3-Hilbert \\ & spaces & spaces & spaces \\ \hline $k = 1$ & H*-algebras & 2-H*-algebras & 3-H*-algebras \\ \hline $k = 2$ &commutative& braided & braided \\ & H*-algebras & 2-H*-algebras & 3-H*-algebras \\ \hline $k = 3$ &`' & symmetric & weakly involutory \\ & & 2-H*-algebras & 3-H*-algebras \\ \hline $k = 4$ &`' & `' &strongly involutory\\ & & & 3-H*-algebras \\ \hline $k = 5$ &`' &`' & `' \\ & & & \\ & & & \\ \hline \end{tabular}} \vskip 1em 2. The quantum-theoretic hierarchy: expected results \end{center} \vskip 0.5em We are far from proving general results along these lines! However, in Section 4 we sketch the structure of $2{\rm Hilb}$ as a strongly involutory 3-H*-algebra, and in Section 5 we define 2-H*-algebras, braided 2-H*-algebra, and symmetric 2-H*-algebras, and describe their relationships to 1-tangles in 2, 3, and 4 dimensions, respectively. An exciting fact about the quantum-theoretic hierarchy is that it automatically subsumes various branches of representation theory. 2-H*-algebras arise naturally as categories of unitary representations of certain Hopf algebras, or more generally `Hopf algebroids', which are to groupoids as Hopf algebras are to groups \cite{Lu}. Braided 2-H*-algebras arise in a similar way from certain quasitriangular Hopf algebroids --- for example, quantum groups --- while symmetric 2-H*-algebras arise from certain triangular Hopf algebroids --- for example, groups. In Section \ref{recon} of this paper we concentrate on the symmetric case. Generalizing the Doplicher-Roberts theorem \cite{DR}, we prove that all symmetric 2-H*-algebras are equivalent to categories of representations of `compact supergroupoids'. If a symmetric 2-H*-algebra is `purely bosonic', it is equivalent to a category of representations of a compact groupoid; if it is `connected', it is equivalent to a category of representations of a compact supergroup. In particular, any connected even symmetric 2-H*-algebra is equivalent to the category ${\rm Rep}(G)$ of continuous unitary finite-dimensional representations of a compact group $G$. This is the original Doplicher-Roberts theorem. One can view our generalized Doplicher-Roberts theorem as a categorified version of the Gelfand-Naimark theorem. The Gelfand-Naimark theorem applies to commutative C*-algebras, but one can easily deduce a version for commutative H*-algebras. Roughly speaking, this says that every commutative H*-algebra $H$ is isomorphic to a commutative H*-algebra of {\it functions} from some set ${\rm Spec}(H)$ to ${\Bbb C}$. Similarly, our theorem implies that every even symmetric 2-H*-algebra $H$ is equivalent to a symmetric 2-H*-algebra of {\it functors} from some groupoid ${\rm Spec}(H)$ to ${\rm Hilb}$. The equivalence is given explicitly by a categorified version of the Gelfand transform. We also construct a categorified version of the Fourier transform, applicable to the representation theory of compact abelian groups. These links between the quantum-theoretic hierarchy and representation theory give new insight into the representation theory of classical groups. The designation of a group as `classical' is more a matter of tradition than of some conceptual definition, but in practice what makes a group `classical' is that it has a nice right universal property. In other words, there is a simple description of homomorphisms into it. Using the fact that group homomorphisms from $G$ to $H$ determine symmetric 2-H*-algebra homomorphisms from ${\rm Rep}(H)$ to ${\rm Rep}(G)$, one can show that for a classical group $H$ the symmetric 2-H*-algebra ${\rm Rep}(H)$ has nice left universal property: there is a simple description of homomorphisms out of it. For example, the group ${\rm U}(n)$ has a distinguished $n$-dimensional unitary representation $\rho$, its fundamental representation on ${\Bbb C}^n$. An $n$-dimensional unitary representation of any group $G$ is essentially the same as a homomorphism from $G$ to ${\rm U}(n)$. Using this right universal property of ${\rm U}(n)$, we show in Section \ref{recon} that the category of unitary representations of ${\rm U}(n)$ is the `free symmetric 2-H*-algebra on one object of dimension $n$'. This statement tersely encodes the usual description of the representations of ${\rm U}(n)$ in terms of Young diagrams. We also give similar characterizations of the categories of representations of other classical groups. In what follows, we denote the composition of 1-morphisms, the horizontal composition of a 1-morphism and a 2-morphism (in either order) and the horizontal composition of 2-morphisms is denoted by $\circ$ or simply juxtaposition. Vertical composition of 2-morphisms is denoted by $\cdot\,$. {\it Nota bene}: in composition we use the ordering in which, for example, the composite of $f \colon x \to y$ and $g \colon y \to z$ is denoted $f\circ g$. We denote the identity morphism of an object $x$ either as $1_x$ or, if there is no danger of confusion, simply as $x$. We refer to our earlier papers on higher-dimensional algebra as HDA0 \cite{BD} and HDA1 \cite{BN}. \section{H*-Categories} Let ${\rm Hilb}$ denote the category whose objects are finite-dimensional Hilbert spaces, and whose morphisms are arbitrary linear maps. (Henceforth, all Hilbert spaces will taken as finite-dimensional unless otherwise specified.) The category ${\rm Hilb}$ is symmetric monoidal, with ${\Bbb C}$ as the unit object, the usual tensor product of Hilbert spaces as the monoidal structure, and the maps \[ S_{x,y}(v \otimes w) = w \otimes v \] as the symmetry, where $x,y \in {\rm Hilb}$, $v \in x$, and $w \in y$. Using enriched category theory \cite{Kelly} we may thus define the notion of a category enriched over ${\rm Hilb}$, or ${\rm Hilb}$-category. Concretely, this amounts to the following: \begin{defn}\hspace{-0.08in}{\bf .}\hspace{0.1in} A ${\rm Hilb}$-{\rm category} $H$ is a category such that for any pair of objects $x,y \in H$ the set of morphisms $\hom(x,y)$ is equipped with the structure of a Hilbert space, and for any objects $x,y,z \in \H$ the composition map \[ \circ\; \colon \hom(x,y) \times \hom(y,z) \to \hom(x,z) \] is bilinear. \end{defn} We may think of the `\hom' in a ${\rm Hilb}$-category $H$ as a functor \[ \hom \colon \H^{{\rm op}} \times H \to {\rm Hilb} \] as follows. An object in $\H^{\rm op} \times H$ is just a pair of objects $(x,y)$ in $\H$, and the $\hom$ functor assigns to this the object $\hom(x,y) \in {\rm Hilb}$. A morphism $F \colon (x,y) \to (x',y')$ in $\H^{\rm op} \times \H$ is just a pair of morphisms $f \colon x' \to x$, $g \colon y \to y'$ in $\H$, and the $\hom$ functor assigns to $F$ the morphism $\hom(F) \colon \hom(x,y) \to \hom(x',y')$ given by \[ \hom(F)(h) = fhg .\] As described in the introduction, we may regard ${\rm Hilb}$ as the categorification of ${\Bbb C}$. A structure on ${\Bbb C}$ which is crucial for Hilbert space theory is complex conjugation, \[ {}^{\overline{\hbox{\hskip 0.5em}}}\, \colon {\Bbb C} \to {\Bbb C}. \] The categorification of this map is a functor \[ {}^{\overline{\hbox{\hskip 0.5em}}}\, \colon {\rm Hilb} \to {\rm Hilb} \] called {\it conjugation}, defined as follows. First, for any Hilbert space $x$, there is a conjugate Hilbert space $\overline x$. This has the same underlying abelian group as $x$, but to keep things straight let us temporarily write $\overline v$ for the element of $\overline x$ corresponding to $v \in x$. Scalar multiplication in $\overline x$ is then given by \[ c\overline v = \overline {(\overline c v)} \] for any $c \in {\Bbb C}$, while the inner product is given by \[ \langle \overline v, \overline w \rangle = \overline {\langle v,w\rangle}. \] Second, for any morphism $f \colon x \to y$ in ${\rm Hilb}$, there is a conjugate morphism $\overline f \colon \overline x \to \overline y$, given by \[ \overline f(\overline v) = \overline{f(v)} \] for all $v \in x$. One can easily check that with these definitions conjugation is a covariant functor. Note that the square of this functor is equal to the identity. Also note that a linear map $f \colon x \to \overline y$ is the same thing as an antilinear (i.e., conjugate-linear) map from $x$ to $y$, while a unitary map $f \colon x \to \overline y$ is the same thing as an antiunitary map from $x$ to $y$. Now, just as in a Hilbert space we have the equation \[ \langle v,w \rangle = \overline {\langle w,v\rangle} \] for any pair of elements, in a 2-Hilbert space we would like an isomorphism \[ \hom(x,y) \cong \overline {\hom(y,x) } \] for every pair of objects. This isomorphism should be be `natural' in some sense, but $\hom(x,y)$ is contravariant in $x$ and covariant in $y$, while $\overline {\hom(y,x)}$ is covariant in $x$ and contravariant in $y$. Luckily ${\rm Hilb}$ is a $\ast$-category, which allows us to define `antinatural isomorphisms' between covariant functors and contravariant functors from any category to ${\rm Hilb}$. This works as follows. In general, a {\it $\ast$-structure} for a category $C$ is defined as a contravariant functor $\ast \colon C \to C$ which acts as the identity on the objects of $C$ and satisfies $\ast^2 = 1_C$. A {\it $\ast$-category} is a category equipped with a $\ast$-structure. For example, ${\rm Hilb}$ is a $\ast$-category where for any morphism $f\colon x \to y$ we define $f^\ast \colon y \to x$ to be the Hilbert space adjoint of $f$: \[ \langle fv, w\rangle = \langle v, f^\ast w \rangle \] for all $v \in x$, $w \in y$. Now suppose that $D$ is a $\ast$-category and $F \colon C \to D$ is a covariant functor, while $G \colon C \to D$ is a contravariant functor. We define an {\it antinatural transformation} $\alpha \colon F \Rightarrow G$ to be a natural transformation from $F$ to $G \circ \ast$. Similarly, an antinatural transformation from $G$ to $F$ is defined to be a natural transformation from $G$ to $F \circ \ast$. As a step towards defining a 2-Hilbert space we now define an H*-category. \begin{defn}\hspace{-0.08in}{\bf .}\hspace{0.1in} An {\rm H*-category} is a ${\rm Hilb}$-category with a $\ast$-structure that defines an antinatural transformation from $\hom(x,y)$ to $\overline {\hom(y,x)}$. \end{defn} \noindent This may require some clarification. Given a $\ast$-structure $\ast \colon \H \to \H$, we obtain for any objects $x,y \in \H$ a function $\ast \colon \hom(x,y) \to \hom(y,x)$. By abuse of notation we may also regard this as a function \[ \ast \colon \hom(x,y) \to \overline{\hom(y,x)}. \] We then demand that this define an antinatural transformation between the covariant functor $\hom \colon H^{\rm op} \times H \to {\rm Hilb}$ to the contravariant functor sending $(x,y) \in H^{\rm op} \times H$ to $\overline{\hom(y,x)} \in {\rm Hilb}$. The following proposition gives a more concrete description of H*-categories: \begin{prop}\label{H*1}\hspace{-0.08in}{\bf .}\hspace{0.1in} An H*-category $\H$ is the same as a ${\rm Hilb}$-category equipped with antilinear maps $\ast \colon \hom(x,y) \to \hom(y,x)$ for all $x,y \in \H$, such that \begin{enumerate} \item $f^{\ast\ast} = f$, \item $(fg)^\ast = g^\ast f^\ast$, \item $\langle fg,h\rangle = \langle g,f^\ast h \rangle$, \item $\langle fg,h\rangle = \langle f, hg^\ast \rangle$ \end{enumerate} whenever both sides of the equation are well-defined. \end{prop} Proof - First suppose that $\H$ is an H*-category. By the antinaturality of $\ast$, for all $x,y \in \H$ there is a linear map $\ast \colon \hom(x,y) \to \overline{\hom(y,x)}$, which is the same as an antilinear map $\ast \colon \hom(x,y) \to \hom(y,x)$. The fact that $\ast$ is a $\ast$-structure implies properties 1 and 2. As for 3 and 4, suppose $(x,y)$ and $(x',y')$ are objects in $\H^{\rm op} \times \H$, and let $(f,g)$ be a morphism from $(x,y)$ to $(x',y')$. The fact that $\ast$ is an antinatural transformation means that the following diagram commutes: \[ \begin{diagram}[\hom(x'y')] \node{\hom(x,y)} \arrow{e,t}{\ast} \arrow{s,l}{V(f,g)} \node{\overline{\hom(y,x)}} \arrow{s,r}{W(f,g)^\ast} \\ \node{\hom(x',y')} \arrow{e,t}{\ast} \node{\overline{\hom(y',x')}} \end{diagram} \] where $V$ is the covariant functor \[ \begin{diagram}[xxxxxxx] \node{\H^{\rm op} \times \H} \arrow{e,t}{\hom} \node{{\rm Hilb}} \end{diagram} \] and $W$ is the contravariant functor \[ \begin{diagram}[xxxxxxx] \node{\H^{\rm op} \times \H} \arrow{e,t}{S_{\H^{\rm op},\H}} \node{\H \times \H^{\rm op}} \arrow{e,t}{\hom} \node{{\rm Hilb}} \arrow{e,t}{{}^{\overline{\hbox{\hskip 0.5em}}}} \node{{\rm Hilb},} \end{diagram} \] where in this latter diagram $S$ denotes the symmetry in ${\rm Cat}$, $\hom$ is regarded as a contravariant functor from $(\H^{\rm op} \times \H)^{\rm op} \cong \H \times \H^{\rm op}$ to ${\rm Hilb}$, and the overline denotes conjugation. This is true if and only if for all $h \in \hom(x,y)$ and $k \in \overline{\hom(y',x')}$, \[ \langle (V(f,g)h)^\ast ,k \rangle = \langle W(f,g)^\ast h^\ast, k\rangle \] or in other words, \[ \langle (fhg)^\ast, k\rangle = \langle h^\ast, gkf\rangle \] or \[ \langle g^\ast h^\ast f^\ast, k\rangle = \langle h^\ast, gkf\rangle .\] Here the inner products are taken in $\overline{\hom(y',x')}$, but the equations also hold with the inner product taken in $\hom(y',x')$. Taking either $f$ or $g$ to be the identity, we obtain 3 and 4 after some relabelling of variables. Conversely, given antilinear maps $\ast \colon \hom(x,y) \to \hom(y,x)$ for all $x,y \in \H$, properties 1 and 2 say that these define a $\ast$-structure for $\H$, and using 3 and 4 we obtain \begin{eqnarray*} \langle g^\ast h^\ast f^\ast , k\rangle &=& \langle g^\ast h^\ast , kf\rangle \\ &=& \langle h^\ast, gkf\rangle ,\end{eqnarray*} showing that $\ast$ is antinatural. \hskip 3em \hbox{\BOX} \vskip 2ex \begin{cor}\label{H*2}\hspace{-0.08in}{\bf .}\hspace{0.1in} If $H$ is an H*-category, for all objects $x,y \in H$ the map $\ast \colon \hom(x,y) \to \hom(y,x)$ is antiunitary. \end{cor} Proof - The map $\ast \colon \hom(x,y) \to \hom(y,x)$ is antilinear, and by 3 and 4 of Proposition \ref{H*1} we have \[ \langle f,g \rangle = \langle g^\ast, f^\ast \rangle = \overline{\langle f^\ast, g^\ast \rangle} \] for all $f,g \in \hom(x,y)$, so $\ast$ is antiunitary. \hskip 3em \hbox{\BOX} \vskip 2ex Next we give a structure theorem for H*-categories. This relies heavily on the theory of `H*-algebras' due to Ambrose \cite{Ambrose}, so let us first recall this theory. For our convenience, we use a somewhat different definition of H*-algebra than that given by Ambrose. Namely, we restrict our attention to finite-dimensional H*-algebras with multiplicative unit, and we do not require the inequality $\|ab\| \le \|a\| \, \|b\|$. \begin{defn} \hspace{-0.08in}{\bf .}\hspace{0.1in} An {\it H*-algebra} $A$ is a Hilbert space that is also an associative algebra with unit, equipped with an antilinear involution $\ast \colon A \to A$ satisfying \begin{eqnarray*} \langle ab,c\rangle &=& \langle b,a^\ast c \rangle \\ \langle ab,c\rangle &=& \langle a, cb^\ast \rangle \end{eqnarray*} for all $a,b,c \in A$. An {\rm isomorphism} of H*-algebras is a unitary operator that is also an involution-preserving algebra isomorphism. \end{defn} The basic example of an H*-algebra is the space of linear operators on a Hilbert space $H$. Here the product is the usual product of operators, the involution is the usual adjoint of operators, and the inner product is given by \[ \langle a,b\rangle = k \,{\rm tr}(a^\ast b) \] where $k > 0$. We denote this H*-algebra by $L^2(H,k)$. It follows from the work of Ambrose that all H*-algebras can be built out of H*-algebras of this form. More precisely, every H*-algebra $A$ is the orthogonal direct sum of finitely many minimal 2-sided ideals $I_i$, each of which is isomorphic as an H*-algebra to $L^2(H_i,k_i)$ for some Hilbert space $H_i$ and some positive real number $k_i$. This result immediately classifies H*-categories with one object. Given an H*-category with one object $x$, ${\rm end}(x)$ is an H*-algebra, and is thus of the above form. Conversely, any H*-algebra is isomorphic to ${\rm end}(x)$ for some H*-category with one object $x$. We generalize this to arbitrary H*-categories as follows. Suppose first that $H$ is an H*-category with finitely many objects. Let $A$ denote the orthogonal direct sum \[ A = \bigoplus_{x,y} \hom(x,y) .\] Then $A$ becomes an H*-algebra if we define the product in $A$ of morphisms in $H$ to be their composite when the composite exists, and zero otherwise, and define the involution in $A$ using the $\ast$-structure of $H$. $A$ is thus the orthogonal direct sum of finitely many minimal 2-sided ideals: \[ A = \bigoplus_{i = 1}^n L^2(H_i, k_i) .\] For each object $x \in H$, the identity morphism $1_x$ can be regarded as an element of $A$. This element is a self-adjoint projection, meaning that \[ 1_x^\ast = 1_x, \qquad 1_x^2 = 1_x .\] It follows that we may write \[ 1_x = \bigoplus_{i = 1}^n p_i^x \] where $p_i^x \in L^2(H_i, k_i)$ is the projection onto some subspace $H_i^x \subseteq H_i$. Note that the elements $1_x$, $x \in H$, form a complete orthogonal set of projections in $A$. In other words, $1_x 1_y = 0$ if $x \ne y$, and \[ \sum_{x \in H} 1_x = 1.\] Thus each Hilbert space $H_i$ is the orthogonal direct sum of the subspaces $H_i^x$. This gives the following structure theorem for H*-categories: \begin{thm}\label{H*3}\hspace{-0.08in}{\bf .}\hspace{0.1in} Let $H$ be an H*-category and $S$ any finite set of objects of $H$. Then for some $n$, there exist positive numbers $k_i > 0$ and Hilbert spaces $H_i^x$ for $i = 1, \dots, n$ and $x \in S$, such that the following hold: \begin{enumerate} \item {\rm For $i = 1, \dots, n$, let \[ H_i = \bigoplus_{x \in S} H_i^x \] denote the orthogonal direct sum, and let $p_i^x$ be the self-adjoint projection from $H_i$ to $H_i^x$. Then for any objects $x,y \in S$, there is a unitary isomorphism between the Hilbert space $\hom(x,y)$ and the subspace \[ \bigoplus_i p_i^x L^2(H_i,k_i) p_i^y \; \subseteq \; \bigoplus_i L^2(H_i, k_i) ,\] Thus we may write any morphism $f \colon x \to y$ as \[ f = \bigoplus_i f_i \] where $f_i \colon H_i^x \to H_i^y$. \item Via the above isomorphism, the composition map \[ \circ\;\colon \hom(x,y) \times \hom(y,z) \to \hom(x,z) \] is given by \[ f \circ g = \bigoplus_i f_i g_i. \] \item Via the same isomorphism, the $\ast$-structure \[ \ast \colon \hom(x,y) \to \hom(y,x) \] is given by \[ f^\ast = \bigoplus_i f_i^\ast .\] \rm} \end{enumerate} \noindent Conversely, given a ${\rm Hilb}$-category $H$ with $\ast$-structure such that the above holds for any finite subset $S$ of its objects, $H$ is an H*-category. \end{thm} Proof - If $H$ has finitely many objects and we take $S$ to be the set of all objects of $H$, properties 1-3 follow from the remarks preceding the theorem. More generally, by Proposition \ref{H*1} any full subcategory of an H*-category is an H*-category, so 1-3 hold for any finite subset $S$ of the objects of $H$. Conversely, given a ${\rm Hilb}$-category $H$ with a $\ast$-structure, if every full subcategory of $H$ with finitely many objects is an H*-category, then $H$ itself is an H*-category. One may check using Proposition \ref{H*1} that if $S$ is any finite subset of the objects of $H$, properties 1-3 imply the full subcategory of $H$ with $S$ as its set of objects is an H*-category. Thus $H$ is an H*-category. \hskip 3em \hbox{\BOX} \vskip 2ex The notions of unitarity and self-adjointness will be important in all that follows. \begin{defn} \hspace{-0.08in}{\bf .}\hspace{0.1in} Let $x$ and $y$ be objects of a $\ast$-category. A morphism $u \colon x \to y$ is {\rm unitary} if $uu^\ast = 1_x$ and $u^\ast u = 1_y$. A morphism $a \colon x \to x$ is {\rm self-adjoint} if $a^\ast = a$. \end{defn} Note that every unitary morphism is an isomorphism. Conversely, the following proposition implies that in an H*-category, isomorphic objects are isomorphic by a unitary. \begin{prop} \label{H*4} \hspace{-0.08in}{\bf .}\hspace{0.1in} Suppose $f \colon x \to y$ is an isomorphism in the H*-category $H$. Then $f = au$ where $a \colon x \to x$ is self-adjoint and $u \colon x \to y$ is unitary. \end{prop} Proof - Suppose that $f \colon x \to y$ is an isomorphism. Then applying Theorem \ref{H*3} to the full subcategory of $H$ with $x$ and $y$ as its only objects, we have $f = \bigoplus f_i$ with $f_i \colon H_i^x \to H_i^y$ an isomorphism for all $i$. Using the polar decomposition theorem we may write $f_i = a_i u_i$, where $a_i \colon H_i^x \to H_i^x$ is the positive square root of $f_i {f_i}^\ast$, and $u_i \colon H_i^x \to H_i^y$ is a unitary operator given by $u_i = a_i^{-1} f_i$. Then defining $a = \bigoplus a_i$ and $u = \bigoplus u_i$, we have $f = au$ where $a$ is self-adjoint and $u$ is unitary. \hskip 3em \hbox{\BOX} \vskip 2ex One can prove a more general polar decomposition theorem allowing one to write any morphism $f \colon x \to y$ in an H*-category as the product of a self-adjoint morphism $a \colon x \to x$ and a partial isometry $i \colon x \to y$, that is, a morphism for which $ii^\ast$ and $i^\ast i$ are self-adjoint idempotents. However, we will not need this result here. \section{2-Hilbert Spaces} The notion of 2-Hilbert space is intended to be the categorification of the notion of Hilbert space. As such, it should be a category having a zero object, direct sums and `direct differences' of objects, tensor products of Hilbert spaces with objects, and `inner products' of objects. So far, with our definition of H*-category, we have formalized the notion of a category in which the `inner product' $\hom(x,y)$ of any two objects $x$ and $y$ is a Hilbert space. Now we deal with the rest of the properties: \begin{defn}\hspace{-0.08in}{\bf .}\hspace{0.1in} A {\rm 2-Hilbert space} is an abelian H*-category. \end{defn} Recall that an abelian category is an ${\rm Ab}$-category (a category enriched over the category ${\rm Ab}$ of abelian groups) such that \begin{enumerate} \item There exists an initial and terminal object. \item Any pair of objects has a biproduct. \item Every morphism has a kernel and cokernel. \item Every monomorphism is a kernel, and every epimorphism is a cokernel. \end{enumerate} Let us comment a bit on what this amounts to. Since an H*-category is enriched over ${\rm Hilb}$ it is automatically enriched over ${\rm Ab}$. We call an initial and terminal object a {\it zero object}, and denote it by $0$. The zero object in a 2-Hilbert space is the analog of the zero vector in a Hilbert space. We call the biproduct of $x$ and $y$ the {\it direct sum}, and denote it by $x \oplus y$. Recall that by definition, this is equipped with morphisms $p_x \colon x \oplus y \to x$, $p_x \colon x \oplus y \to y$, $i_x \colon x \to x \oplus y$, $i_y \colon y \to x \oplus y$ such that \[ i_x p_x = 1_x , \qquad i_y p_y = 1_y, \qquad p_x i_x + p_y i_y = 1_{x\oplus y} .\] The direct sums in a 2-Hilbert space are the analog of addition in a Hilbert space. Similarly, the cokernels in a 2-Hilbert space are the analogs of differences in a Hilbert space. Finally, the ability to tensor objects in a 2-Hilbert space by Hilbert spaces (the analog of scalar multiplication) will follow from the other properties, so we do not need to include it in the definition of 2-Hilbert space. Some aspects of our definition of 2-Hilbert space may seem unmotivated by the analogy with Hilbert spaces. Why should a 2-Hilbert space have kernels, and why should it satisfy clause 4 in the definition of abelian category? In fact, these properties follow from the rest. \begin{prop}\hspace{-0.08in}{\bf .}\hspace{0.1in} \label{2hilb1} Let $H$ be an H*-category. Then the following are equivalent: \begin{enumerate} \item There exists an initial object. \item There exists a terminal object. \item There exists a zero object. \end{enumerate} Moreover, the following are equivalent: \begin{enumerate} \item Every pair of objects has a product. \item Every pair of objects has a coproduct. \item Every pair of objects has a direct sum. \end{enumerate} Moreover, the following are equivalent: \begin{enumerate} \item Every morphism has a kernel. \item Every morphism has a cokernel. \end{enumerate} Finally, if $H$ has a zero object, every pair of objects in $H$ has a direct sum, and every morphism in $H$ has a cokernel, then $H$ is a semisimple abelian category. \end{prop} Proof - It is well-known \cite{MacLane} that an initial or terminal object in an ${\rm Ab}$-category is automatically a zero object. Alternatively, this is true in every $\ast$-category, using the bijection $\ast \colon \hom(x,y) \to \hom(y,x)$. It is also well-known that in an ${\rm Ab}$-category, a binary product or coproduct is automatically a binary biproduct. Furthermore, it is easy to check that in any $\ast$-category, the morphism $j \colon k \to x$ is a kernel of $f\colon x \to y$ if and only if $j^\ast \colon x \to k$ is a cokernel of $f^\ast \colon y \to x$. Thus a $\ast$-category has kernels if and only if it has cokernels. Now suppose that $H$ is an H*-category with a zero object, direct sums, and cokernels. Then $H$ has kernels as well, so to show $H$ is abelian we merely need to prove that every monomorphism is a kernel and every epimorphism is a cokernel. Let us show a monomorphism $f \colon x \to y$ is a kernel; it follows using the $\ast$-structure that every epimorphism is a cokernel. It suffices to show this result for any full subcategory of $H$ with finitely many objects, so by Theorem \ref{H*3} we may write \[ f = \bigoplus_i f_i \] where $f_i \colon H_i^x \to H_i^y$ is a linear operator. Let $p \colon y \to y$ be given by $\bigoplus p_i$ where $p_i$ is the projection onto the orthogonal complement of the range of $f_i$. We claim that $f\colon x \to y$ is a kernel of $p$. Since $f_i p_i = 0$ for all $i$ we have $fp = 0$. We also need to show that if $f' \colon x' \to y$ is any morphism with $f'p = 0$, then there is a unique $g \colon x' \to x$ with $f' = gf$. Writing $f' = \bigoplus f'_i$, the fact that $f'p = 0$ implies that the range of $f'_i$ is contained in the range of $f_i$. Thus by linear algebra there exists $g_i \colon H_i^{x'} \to H_i^x$ such that $f'_i = g_i f_i$. Letting $g = \bigoplus g_i$, we have $f' = gf$, and $g$ is unique with this property because $f$ is monic. Finally, note that $H$ is semisimple, i.e., every short exact sequence splits. This follows from Theorem \ref{H*3} and elementary linear algebra. \hskip 3em \hbox{\BOX} \vskip 2ex Given a 2-Hilbert space $H$, the fact that $H$ is semisimple implies that every object is isomorphic to a direct sum of {\it simple} objects, that is, objects $x$ for which ${\rm end}(x)$ is isomorphic as an algebra to ${\Bbb C}$. This fact lets us reason about 2-Hilbert spaces using bases: \begin{defn} \hspace{-0.08in}{\bf .}\hspace{0.1in} Given a 2-Hilbert space $H$, a set of nonisomorphic simple objects of $H$ is called a {\rm basis} if every object of $H$ is isomorphic to a finite direct sum of objects in that set. \end{defn} \begin{cor} \label{2hilb2} \hspace{-0.08in}{\bf .}\hspace{0.1in} Every 2-Hilbert space $H$ has a basis, and any two bases of $H$ have the same cardinality. \end{cor} Proof - The 2-Hilbert space $H$ has a basis because it is semisimple: given any Given two bases $\{e_\alpha\}$ and $\{f_\beta\}$, each object $e_\alpha$ is isomorphic to a direct sum of copies of the objects $e_\beta$, but as the $e_\alpha$ and $f_\beta$ are simple we must actually have an isomorphism $e_\alpha \cong f_\beta$ for some $\beta$. This $\beta$ is unique since no distinct $f_\beta$'s are isomorphic. This sets up a function from $\{e_\alpha\}$ to $\{f_\beta\}$, and similar reasoning gives us the inverse function. \hskip 3em \hbox{\BOX} \vskip 2ex \begin{defn} \hspace{-0.08in}{\bf .}\hspace{0.1in} The {\rm dimension} of a 2-Hilbert space is the cardinality of any basis of it. \end{defn} Note that every basis $\{e_\alpha\}$ of a 2-Hilbert space is `orthogonal' in the sense that \[ \hom(e_\alpha, e_\beta) \cong \cases{L^2({\Bbb C},k_\alpha) & $\alpha = \beta$ \cr 0 & $\alpha \ne \beta$ } \] where the isomorphism is one of H*-algebras, and $k_\alpha$ are certain positive constants. Moreover, up to reordering, the constants $k_\alpha$ are independent of the choice of basis. For suppose $x,y$ are two isomorphic objects in an H*-category. By Proposition \ref{H*4} there is a unitary isomorphism $f \colon x \to y$. Then there is an H*-algebra isomorphism $\alpha \colon {\rm end}(x) \to {\rm end}(y)$ given by $\alpha(g) = f^{-1}gf$. One would also like to be able to tensor objects in a 2-Hilbert space with Hilbert spaces, but this is a consequence of the definition we have given, since one may define the tensor product of an object $x$ in a 2-Hilbert space with an $n$-dimensional Hilbert space to be the direct sum of $n$ copies of $x$. In fact, ${\rm Hilb}$ has a structure analogous to that of an algebra, with tensor product and direct sum playing the roles of multiplication and addition. In the terminology we introduce in Section \ref{2-H*-algebras}, one says that ${\rm Hilb}$ is a `2-H*-algebra'. One can develop a theory of modules of 2-H*-algebras following the ideas of Kapranov and Voevodsky \cite{KV} and Yetter \cite{Yetter}. Every 2-Hilbert space $H$ is then a module over ${\rm Hilb}$. We will not pursue this further here. \section{2Hilb as a 2-Category}\label{2Hilb} We now investigate a certain 2-category $2{\rm Hilb}$ of 2-Hilbert spaces. To keep things simple we take as its objects only finite-dimensional 2-Hilbert spaces. Nonetheless we prove theorems more generally whenever possible. \begin{defn} \hspace{-0.08in}{\bf .}\hspace{0.1in} A {\rm morphism} $F \colon H \to H'$ between 2-Hilbert spaces $H$ and $H'$ is an exact functor such that $F \colon \hom(x,y) \to \hom(F(x), F(y))$ is linear and $F(f^\ast) = F(f)^\ast$ for all $f \in \hom(x,y)$. \end{defn} Recall that an exact functor is one preserving short exact sequences. Exactness is an natural sort of condition for functors between abelian categories. Similarly, the requirement that $F\colon \hom(x,y) \to \hom(F(x), F(y))$ be linear is a natural condition for functors between ${\rm Hilb}$-categories; one calls such a functor a {\it ${\rm Hilb}$-functor}. Finally, $F(f^\ast) = F(f)^\ast$ is a natural condition for functors between $\ast$-categories, and functors satisfying it are called {\it $\ast$-functors}. The following fact is occaisionally handy: \begin{prop} \hspace{-0.08in}{\bf .}\hspace{0.1in} Let $F \colon H \to H'$ be a functor between 2-Hilbert spaces such that for all $x,y \in H$, $F \colon \hom(x,y) \to \hom(F(x), F(y))$ is linear. Then the following are equivalent: \begin{enumerate} \item $F$ is exact. \item $F$ is left exact. \item $F$ is right exact. \item $F$ preserves direct sums. \end{enumerate} \end{prop} Proof - Following Yetter \cite{Yetter}, we use the fact that every short exact sequence splits. \hbox{\hskip 30em} \hskip 3em \hbox{\BOX} \vskip 2ex \begin{defn} \hspace{-0.08in}{\bf .}\hspace{0.1in} A {\rm 2-morphism} $\alpha \colon F \Rightarrow F'$ between morphisms $F,F' \colon H \to H'$ between 2-Hilbert spaces $H$ and $H'$ is a natural transformation. \end{defn} \begin{defn} \hspace{-0.08in}{\bf .}\hspace{0.1in} We define the 2-category $2{\rm Hilb}$ to be that for which objects are finite-dimensional 2-Hilbert spaces, while morphisms and 2-morphisms are defined as above. \end{defn} Now, just as in some sense ${\Bbb C}$ is the primordial Hilbert space and ${\rm Hilb}$ is the primordial 2-Hilbert space, $2{\rm Hilb}$ should be the primordial 3-Hilbert space. The study of $2{\rm Hilb}$ should thus shed light on the properties of the still poorly understood 3-Hilbert spaces. However, note that ${\Bbb C}$ is not merely a Hilbert space, but also a commutative monoid, in fact a commutative H*-algebra. Similarly, ${\rm Hilb}$ is not merely a 2-Hilbert space, but also a symmetric monoidal category when equipped with its usual tensor product. Indeed, in Section \ref{2-H*-algebras} we show that ${\rm Hilb}$ is a `symmetric 2-H*-algebra'. Likewise, we expect $2{\rm Hilb}$ to be not only a 3-Hilbert space, but also a strongly involutory monoidal 2-category, in fact a `strongly involutory 3-H*-algebra'. As sketched in HDA0, commutative monoids, symmetric monoidal categories, and strongly involutory monoidal 2-categories are all examples of `stable' $n$-categories. In general we expect $n{\rm Hilb}$ to be a `stable $(n+1)$-H*-algebra.' The results below offer some support for this expectation. We begin with a study of duality in $2{\rm Hilb}$, as this is the most distinctive aspect of Hilbert space theory. Note that every element $x \in {\Bbb C}$ has a kind of `dual' element, namely, its complex conjugate $\overline x$. Similarly, the category ${\rm Hilb}$ has duality both for objects and for morphisms. At the level of morphisms, each linear map $f \colon x \to y$ between Hilbert spaces has a dual $f^\ast \colon y \to x$, the usual Hilbert space adjoint of $f$. This defines a $\ast$-structure on $H$. Duality at the level of objects can be regarded either as a contravariant functor assigning to each each Hilbert space $x$ its dual $x^\ast$, or as a covariant functor assigning to each Hilbert space $x$ its conjugate $\overline x$. These two viewpoints become equivalent if we take advantage of duality at the morphism level, since $x^\ast$ and $\overline x$ are antinaturally isomorphic. Similarly, $2{\rm Hilb}$ has duality for objects, morphisms, and 2-morphisms. As in ${\rm Hilb}$, we can use duality at a given level to reinterpret dualities at lower levels in various ways. This recursive process can become rather confusing unless we choose by convention to take certain dualities as `basic' and others as derived. Here we follow the philosophy of HDA0: any 2-morphism $\alpha \colon F \Rightarrow G$ has a dual $\alpha^\ast \colon G \Rightarrow F$, any morphism $F \colon H \to H'$ has a dual $F^\ast \colon H' \to H$, and every object $H$ has a dual $H^\ast$. (Our notation differs from HDA0 in that we use the same symbol to denote all these different levels of duality.) \subsection{Duality for 2-morphisms} Duals of 2-morphisms are the easiest to define. It pays to do so in the greatest possible generality: \begin{defn} \label{star} \hspace{-0.08in}{\bf .}\hspace{0.1in} Given a category $C$ and a $\ast$-category $D$, the {\rm dual} $\alpha^\ast$ of a natural transformation $\alpha \colon F \Rightarrow G$ is the natural transformation with $(\alpha^\ast)_c = (\alpha_c)^\ast$ for all $c \in C$. \end{defn} It is easy to check that $\alpha^\ast$ is a natural transformation when $\alpha$ is, and that \[ (\alpha^\ast)^\ast = \alpha, \qquad 1^\ast = 1.\] The vertical composite of natural transformations satisfies \[ (\alpha\cdot \beta)^\ast = \beta^\ast\cdot \alpha^\ast \] when this is defined. When $D$ is a $\ast$-category, the horizontal composite of a functor $F \colon B \to C$ and a natural transformation $\alpha \colon G \Rightarrow H$ with $G,H \colon C \to D$ satisfies \[ (F\alpha)^\ast = F \alpha^\ast .\] Similarly, when $F \colon C \to D$ is a $\ast$-functor and $\alpha \colon G \Rightarrow H$ is a natural transformation between $G,H \colon B \to C$, we have \[ (\alpha F)^\ast = \alpha^\ast F.\] In particular, taking $C,D$ to be 2-Hilbert spaces, we obtain the definition of the dual of a 2-morphism in $2{\rm Hilb}$. We also obtain the notion of `unitary' and `self-adjoint' natural transformations: \begin{defn} \hspace{-0.08in}{\bf .}\hspace{0.1in} Given a category $C$, a $\ast$-category $D$, and functors $F,G \colon C \to D$, a natural transformation $\alpha \colon F \Rightarrow G$ is {\rm unitary} if \[ \alpha \alpha^\ast = 1_F , \qquad \alpha^\ast \alpha = 1_G .\] A natural transformation $\alpha \colon F \Rightarrow F$ is {\rm self-adjoint} if \[ \alpha^\ast = \alpha .\] \end{defn} Equivalently, $\alpha$ is unitary if $\alpha_c$ is a unitary morphism in $D$ for all objects $c \in C$, and self-adjoint if $\alpha_c$ is self-adjoint for all $c \in C$. Note that every unitary natural transformation is a natural isomorphism. Conversely: \begin{prop} \label{unitary.nat} \hspace{-0.08in}{\bf .}\hspace{0.1in} Suppose $F,G \colon H \to H'$ are morphisms between 2-Hilbert spaces and $\alpha \colon F \Rightarrow G$ is a natural isomorphism. Then $\alpha = \beta\cdot\gamma$ where $\beta \colon F \Rightarrow F$ is self-adjoint and $\gamma \colon F \Rightarrow G$ is unitary. \end{prop} Proof - By Proposition \ref{H*4}, for any $x \in H$ we can write the isomorphism $\alpha_x \colon F(x) \to G(x)$ as the composite $\beta_x \gamma_x$, where $\beta_x \colon F(x) \to F(x)$ is self-adjoint and $\gamma_x \colon F(x) \to G(x)$ is unitary. More importantly, the polar decomposition gives a natural way to construct $\beta_x$ and $\gamma_x$ from $\alpha_x$: we take $\beta_x$ to be the positive square root of $\alpha_x {\alpha_x}^\ast$, and take $\gamma_x = \beta^{-1}_x \alpha_x$. Since $\alpha \alpha^\ast$ is a natural transformation from $F$ to itself, if we define $P(\alpha \alpha^\ast)_x = P(\alpha_x {\alpha_x}^\ast)$ for any polynomial $P$, we have \[ P(\alpha_x {\alpha_x}^\ast) F(f) = F(f) P(\alpha_y {\alpha_y}^\ast) \] for any morphism $f \colon x \to y$. By the finite-dimensional spectral theorem, we can find a sequence of polynomials $P_i$ such that $P_i(\alpha_x {\alpha_x}^\ast) \to \beta_x$ and $P_i(\alpha_y {\alpha_y}^\ast) \to \beta_y$. Thus \[ \beta_x F(f) = F(f) \beta_y , \] so $\beta$ is a natural transformation from $F$ to itself. It follows that $\gamma = \beta^{-1}\cdot \alpha$ is a natural transformation from $F$ to $G$. Clearly $\beta$ is self-adjoint and $\gamma$ is unitary. \hskip 3em \hbox{\BOX} \vskip 2ex \subsection{Duality for morphisms} Duals of morphisms in $2{\rm Hilb}$ are just adjoint functors. Normally one needs to distinguish between left and right adjoint functors, but duality at the 2-morphism level allows us to turn left adjoints into right adjoints, and vice versa: \begin{prop}\hspace{-0.08in}{\bf .}\hspace{0.1in} \label{2cat1} Suppose $F \colon H \to H'$, $G \colon H' \to H$ are morphisms in $2{\rm Hilb}$. Then $F$ is left adjoint to $G$ with unit $\iota \colon 1_H \Rightarrow FG$ and counit $\epsilon \colon GF \Rightarrow 1_{H'}$ if and only if $F$ is right adjoint to $G$ with unit $\epsilon^\ast \colon 1_{H'} \Rightarrow GF$ and counit $\iota^\ast \colon FG \Rightarrow 1_H$. \end{prop} Proof - The triangle equations for $\iota$ and $\epsilon$: \[ (\iota F) \cdot (F \epsilon) = 1_F, \qquad (G \iota) \cdot (\epsilon G) = 1_G ,\] become equivalent to those for $\epsilon^\ast$ and $\iota^\ast$: \[ (\epsilon^\ast G) \cdot (G \iota^\ast) = 1_G, \qquad (F \epsilon^\ast) \cdot (\iota^\ast F) = 1_F, \] by taking duals. \hskip 3em \hbox{\BOX} \vskip 2ex \noindent As noted by Dolan \cite{Dolan}, it is probably quite generally true in $n$-categories that duality for $j$-morphisms allows us to turn `left duals' of $(j-1)$-morphisms into `right duals' and vice versa. This should give the theory of $n$-Hilbert spaces quite a different flavor from general $n$-category theory. Every morphism in $2{\rm Hilb}$ has an adjoint. We prove this using bases and the concept of a skeletal 2-Hilbert space. \begin{defn}\hspace{-0.08in}{\bf .}\hspace{0.1in} A category is {\rm skeletal} if all isomorphic objects are equal. \end{defn} \begin{defn}\hspace{-0.08in}{\bf .}\hspace{0.1in} A {\rm unitary equivalence} between 2-Hilbert spaces $H$ and $H'$ consists of morphisms $U \colon H \to H'$, $V \colon H' \to H$ and unitary natural transformations $\iota \colon 1_H \Rightarrow UV$, $\epsilon \colon VU \Rightarrow 1_{H'}$ forming an adjunction. If there exists a unitary equivalence between $H$ and $H'$, we say they are {\rm unitarily equivalent}. \end{defn} \begin{prop}\hspace{-0.08in}{\bf .}\hspace{0.1in} \label{2cat2} Any 2-Hilbert space is unitarily equivalent to a skeletal 2-Hilbert space. \end{prop} Proof - Let $\{e_\lambda\}$ be a basis for the 2-Hilbert space $H$. For any nonnegative integers $\{n^\lambda\}$ with only finitely many nonzero, make a choice of direct sum \[ \bigoplus_\lambda n^\lambda e_\lambda ,\] where $n^\lambda e_\lambda$ denotes the direct sum of $n^\lambda$ copies of $e_\lambda$. (Recall that the direct sum is an object equipped with particular morphisms; it is only unique up to isomorphism, but here we fix a particular choice.) Let $H_0$ denote the full subcategory of $H$ with only these direct sums as objects. Note that $H_0$ inherits a 2-Hilbert space structure from $H$, and it is skeletal. Let $V \colon H_0 \to H$ denote the inclusion functor. For any $x \in H$ there is a unique object $U(x) \in H_0$ for which $V(U(x))$ is isomorphic to $x$. By Proposition \ref{H*4}, we may choose a unitary isomorphism \[ \iota_x \colon x \to V(U(x)). \] For $x = V(y)$ we have $U(x) = y$, so we choose $\iota_x$ to be the identity in this case. For each morphism $f\colon x \to y$ define $U(f) \colon U(x) \to U(x')$ so that the following diagram commutes: \[ \begin{diagram}[V(U(x))] \node{x} \arrow{e,t}{f} \arrow{s,l}{\iota_x} \node{y} \arrow{s,r}{\iota_{y}} \\ \node{V(U(x))} \arrow{e,b}{V(U(f))} \node{V(U(y))} \end{diagram} \] It follows that $U \colon H \to H_0$ is a functor. One may check that $U$ and $V$ are actually morphisms of 2-Hilbert spaces. Moreover, one may check that there is a natural isomorphism \[ \hom(Ux,y) \cong \hom(x,Vy) \] given by \[ f \mapsto \iota_x V(f). \] It follows that $U$ is left adjoint to $V$. The unit of this adjunction is $\iota$, while the counit is the identity. These are both unitary natural transformations. \hskip 3em \hbox{\BOX} \vskip 2ex Just as with Hilbert spaces, phrasing definitions and theorems about 2-Hilbert spaces in terms of a basis is usually a mistake, since they should be manifestly invariant under unitary equivalence. In comparison, the use of bases to prove theorems is at worst a minor lapse of taste, and sometimes convenient. This is facilitated by the use of skeletal 2-Hilbert spaces. \begin{prop}\hspace{-0.08in}{\bf .}\hspace{0.1in} \label{2cat3} Let $F \colon H \to H'$ be a morphism in $2{\rm Hilb}$. Then there is a morphism $F^\ast \colon H' \to H$ that is left and right adjoint to $F$. \end{prop} Proof - Here we opt for a lowbrow proof using bases, to illustrate the analogy between an adjoint functor and the adjoint of a matrix. By Proposition \ref{2cat2} it suffices to consider the case where $H$ and $H'$ are skeletal. Let $\{e_\lambda\}$ be a basis for $H$ and $\{e'_\mu\}$ a basis for $H'$. Write \[ F(e_\lambda) = \bigoplus_\mu F_{\lambda \mu} e'_\mu \] where $F_{\lambda\mu}$ are nonnegative integers and $F_{\lambda\mu} e'_\mu$ denotes the direct sum of $F_{\lambda\mu}$ copies of $e'_\mu$. Let \[ {F_{\mu\lambda}}^\ast = F_{\lambda\mu}. \] Defining \[ F^\ast(e'_\mu) = \bigoplus_{\mu} F^\ast_{\mu\lambda} e_\lambda ,\] one may check that $F^\ast$ extends uniquely to a morphism from $H'$ to $H$. Note that both $\hom(F e_\lambda, e'_\mu)$ and $\hom(e_\lambda, F^\ast e'_\mu)$ may be naturally identified with a direct sum of $F_{\lambda\mu}$ copies of ${\Bbb C}$, which sets up an isomorphism $\hom(F e_\lambda, e'_\mu) \cong \hom(e_\lambda, F^\ast e'_\mu)$. One can check that this extends uniquely to a natural isomorphism \[ \hom(Fx,y) \cong \hom(x,F^\ast y), \] so $F^\ast$ is a right adjoint, and by Proposition \ref{2cat1} also a left adjoint, of $F$. \hskip 3em \hbox{\BOX} \vskip 2ex A basic fact in Hilbert space theory is that two objects in ${\rm Hilb}$ are isomorphic if and only if there is a unitary morphism between them. The same is true of objects in any other 2-Hilbert space, by Proposition \ref{H*4}. Similarly, two morphisms in $2{\rm Hilb}$ are isomorphic if and only if there is a unitary natural transformation between them, by Proposition \ref{unitary.nat}. Below we show a similar result for objects in $2{\rm Hilb}$. In general, we expect a recursively defined notion of `equivalence' of $j$-morphisms in an $n$-category: two $n$-morphisms are equivalent if they are equal, while two $(j-1)$-morphisms $x,y$ are equivalent if there exist $f \colon x \to y$ and $g \colon y \to x$ with $gf$ and $fg$ equivalent to the identity on $x$ and $y$, respectively. In an $n$-Hilbert space we also expect a similar notion of `unitary equivalence': two $(n-1)$-morphisms are unitarily equivalent if they are equal, while two $(j-1)$-morphisms $x,y$ are unitarily equivalent if there exists $u \colon x \to y$ with $uu^\ast$ and $u^\ast u$ unitarily equivalent to $1_x$ and $1_y$, respectively. Our results so far lead us to suspect that, quite generally, equivalent $j$-morphisms in an $n$-Hilbert space will be unitarily equivalent. \begin{defn}\label{equivalence}\hspace{-0.08in}{\bf .}\hspace{0.1in} An {\rm equivalence} between 2-Hilbert spaces $H$ and $H'$ is an pair of morphisms $F \colon H \to H'$, $G \colon H' \to H$ together with natural isomorphisms $\alpha \colon 1_H \Rightarrow FG$, $\beta \colon GF \Rightarrow 1_{H'}$. If there is an equivalence between $H$ and $H'$, we say they are {\rm equivalent}. \end{defn} Note that a unitary equivalence is automatically an equivalence. Conversely: \begin{prop} \label{unitary.eq} \hspace{-0.08in}{\bf .}\hspace{0.1in} Suppose $H$ and $H'$ are 2-Hilbert spaces and the morphisms $F \colon H \to H'$, $G \colon H' \to H$ can be extended to an equivalence between $H$ and $H'$. Then $F$ and $G$ can be extended to a unitary equivalence between $H$ and $H'$. \end{prop} Proof - Suppose $\alpha \colon 1_H \Rightarrow FG$, $\beta \colon GF \Rightarrow 1_{H'}$ are natural isomorphisms. By Proposition \ref{unitary.nat} we can find unitary natural transformations $\gamma \colon 1_H \Rightarrow FG$, $\delta \colon GF \Rightarrow 1_{H'}$. We may then obtain an adjunction by replacing $\gamma$ with the composite $\gamma'$ given by \[ \begin{diagram} [FG = F1_{H'}G] \node{1_H} \arrow{e,t}{\gamma} \node{FG = F1_{H'}G} \arrow{e,t}{F\delta^{-1}G} \node{FGFG} \arrow{e,t}{\gamma^{-1}FG} \node{FG} \end{diagram} \] Checking that this is an adjunction is a lengthy but straightforward calculation. Noting that $\gamma'$ is unitary, we conclude that $(F,G,\iota,\epsilon)$ is a unitary equivalence. \hskip 3em \hbox{\BOX} \vskip 2ex \noindent When we are being less pedantic, we call a 2-Hilbert space morphism $F \colon H \to H'$ an {\it equivalence} if it can be extended to an equivalence in the sense of Definition \ref{equivalence}. Just as Hilbert spaces are classified by their dimension, we have: \begin{cor} \label{2cat4} \hspace{-0.08in}{\bf .}\hspace{0.1in} Two 2-Hilbert spaces are equivalent if and only if they have the same dimension. \end{cor} Proof - Since an equivalence between $H$ and $H'$ carries a basis of $H$ to a basis of $H'$, Proposition \ref{2hilb2} implies that dimension is preserved by equivalence. By Proposition \ref{2cat2} it thus suffices to show two skeletal 2-Hilbert spaces are equivalent if they have the same dimension. Let $\{e_\lambda\}$ be a basis of $H$ and $\{e'_\lambda\}$ a corresponding basis of $H'$. Then there is a unique 2-Hilbert space morphism with $F(e_\lambda) = e'_\lambda$, and the adjunction constructed as in the proof of Proposition \ref{2cat3} is a unitary equivalence. \hskip 3em \hbox{\BOX} \vskip 2ex \subsection{Duality for objects} \label{dualityforobjects} Finally, duals of objects in $2{\rm Hilb}$ are defined using an `internal hom'. Given 2-Hilbert spaces $H$ and $H'$, let $\hom(H,H')$ be the category having 2-Hilbert space morphisms $F\colon H \to H'$ as objects and 2-morphisms between these as morphisms. \begin{prop} \label{2cat5}\hspace{-0.08in}{\bf .}\hspace{0.1in} Suppose $H$ is a finite-dimensional 2-Hilbert space and $H'$ is a 2-Hilbert space. Then the category $\hom(H,H')$ becomes a ${\rm Hilb}$-category if for any $F,G \in \hom(H,H')$ we make $\hom(F,G)$ into a Hilbert space with the obvious linear structure and the inner product given by \[ \langle \alpha, \beta \rangle = \sum_{\lambda} \langle \alpha_{e_\lambda} , \beta_{e_\lambda} \rangle \] for any basis $\{e_\lambda\}$ of $H$. Moreover, $\hom(H,H')$ becomes a 2-Hilbert space if we define the dual of $\alpha \colon F \Rightarrow G$ by $(\alpha^\ast)_x = (\alpha_x)^\ast$. \end{prop} Proof - Note first that $\hom(F,G)$ becomes a vector space if we define \[ (\alpha + \beta)_x = \alpha_x + \beta_x, \qquad (c\alpha)_x = c(\alpha_x) \] for any $\alpha, \beta \colon F \Rightarrow G$ and $c \in {\Bbb C}$. Note also that the inner product described above is nondegenerate, since if $\alpha_{e_\lambda} = 0$ for all objects $e_\lambda$ in a basis, then $\alpha = 0$. Finally, note that the inner product is independent of the choice of basis: if $\{e'_\lambda\}$ is another basis we may assume after reordering that $e_\lambda \cong e'_\lambda$, and by Proposition \ref{H*4} we may choose unitary isomorphisms $u_\lambda \colon e_\lambda \to e'_\lambda$, so that \[ \alpha_{e'_\lambda} = F(u_\lambda)^\ast \alpha_{e_\lambda} G(u_\lambda) \] and similarly for $\beta$. It follows that \begin{eqnarray*} \langle \alpha_{e'_\lambda} , \beta_{e'_\lambda} \rangle &=& \langle F(u_\lambda)^\ast \alpha_{e_\lambda} G(u_\lambda) , F(u_\lambda)^\ast \beta_{e_\lambda} G(u_\lambda) \rangle \\ &=& \langle \alpha_{e_\lambda} , \beta_{e_\lambda} \rangle .\end{eqnarray*} Since composition of morphisms in $\hom(H,H')$ is bilinear, it becomes a ${\rm Hilb}$-category. It is easy to check that defining $(\alpha^\ast)_x = (\alpha_x)^\ast$ makes $\hom(H,H')$ into a $\ast$-category, and using Proposition \ref{H*1} one can also check it is an H*-category. To check that it is a 2-Hilbert space it suffices by Proposition \ref{2hilb1} to check that it has a zero object, direct sums and kernels. Any functor $0 \colon H \to H'$ mapping all objects in $H$ to zero objects in $H'$ is initial in $\hom(H,H')$. Given $F,F' \in \hom(H,H')$, we may take as the direct sum $F \oplus F'$ any functor with $(F \oplus F')(x) = F(x) \oplus F(x')$ for any object $x \in H$ and $(F \oplus F')(f) = F(f) \oplus F(f')$ for any morphism $f$. Similarly, given $\alpha \colon F \to F'$, we may construct $\ker \alpha \in \hom(H,H')$ by letting $(\ker \alpha)(x) = \ker \alpha_x$ for any object $x$ and defining $(\ker \alpha)(f)$ for any morphism using the universal property of the kernel. \hbox{\hskip 30em} \hskip 3em \hbox{\BOX} \vskip 2ex \begin{defn} \hspace{-0.08in}{\bf .}\hspace{0.1in} Given a finite-dimensional 2-Hilbert space $H$, the {\rm dual} $H^\ast$ is the 2-Hilbert space $\hom(H,{\rm Hilb})$. \end{defn} The following is an analog of the Riesz representation theorem for finite-dimensional 2-Hilbert spaces. In its finite-dimensional form, the Riesz representation theorem says if $x$ is a Hilbert space, any morphism $f \colon x \to {\Bbb C}$ is equal to one of the form \[ \langle v, \cdot \rangle \] for some $v \in H$. This determines an isomorphism $\overline x \cong x^\ast$. Similarly, given a 2-Hilbert space $H$, we say a morphism $F \colon H \to {\rm Hilb}$ is {\it representable} if it is naturally isomorphic to one of the form \[ \hom(x, \cdot) \] for some $x \in H$. The essence of the Riesz representation theorem for 2-Hilbert spaces is that every morphism $F \colon H \to {\rm Hilb}$ is representable. This yields an equivalence between $H^{\rm op}$ and $H^\ast$. \begin{prop} \hspace{-0.08in}{\bf .}\hspace{0.1in} For any finite-dimensional 2-Hilbert space $H$, the morphism \break $U \colon H^{\rm op} \to H^\ast$ given by \[ U(x) = \hom(x,\cdot) , \qquad U(f) = \hom(f,\cdot) \] is an equivalence between $H^{\rm op}$ and $H^\ast$. \end{prop} Proof - It suffices to show that $U$ is fully faithful and essentially surjective. We can check both of these using a basis $\{e_\lambda\}$ of $H$. We leave the full faithfulness to the reader. Checking that $U$ is essentially surjective amounts to checking that any $F \in H^\ast$ is representable. Note there is a `dual basis' of 2-Hilbert space morphisms $f^\lambda \in \hom(H,{\rm Hilb})$ with \[ f^\mu(e_\lambda) \cong \cases{{\Bbb C} & $\lambda = \mu$ \cr 0 & $\lambda \ne \mu$ } \] Since any morphism $F \colon H \to {\rm Hilb}$ is determined up to natural isomorphism by its value on the basis $\{e_\lambda\}$, any $F \in H^\ast$ is isomorphic to a direct sum of the $\{f^\lambda\}$. But $f^\lambda$ is isomorphic to $U(e_\lambda)$, so $U$ is essentially surjective. \hskip 3em \hbox{\BOX} \vskip 2ex \subsection{The tensor product}\label{tensorproduct} Next we develop the tensor product of 2-Hilbert spaces. For this we need the analog of a bilinear map: \begin{defn} \hspace{-0.08in}{\bf .}\hspace{0.1in} Given 2-Hilbert spaces $H,H',K$, a functor $F \colon H \times H' \to K$ is a {\rm bimorphism} of 2-Hilbert spaces if for any objects $x \in H$, $x' \in H'$ the functors $F(x \otimes \,\cdot\,) \colon H' \to K$ and $F(\,\cdot \,\otimes x') \colon H \to K$ are 2-Hilbert space morphisms. We write ${\rm bihom}(H\times H', K)$ for the category having bimorphisms $F \colon H \times H' \to K$ as objects and natural transformations between these as morphisms. \end{defn} \begin{prop} \hspace{-0.08in}{\bf .}\hspace{0.1in} Suppose $H$ and $H'$ are finite-dimensional 2-Hilbert spaces and $K$ is a 2-Hilbert space. Then ${\rm bihom}(H\times H',K)$ becomes a ${\rm Hilb}$-category if for any $F,G \in {\rm bihom}(H \times H',K)$ we make $\hom(F,G)$ into a Hilbert space with the obvious linear structure and the inner product given by \[ \langle \alpha, \beta \rangle = \sum_{\lambda,\mu} \langle \alpha_{(e_\lambda,f_\mu)} , \beta_{(e_\lambda,f_\mu)} \rangle \] for any bases $\{e_\lambda\}$ of $H$ and $\{f_\mu\}$ of $H'$. Moreover, ${\rm bihom}(H \times H',K)$ becomes a 2-Hilbert space if we define the dual of $\alpha \colon F \Rightarrow G$ by $(\alpha^\ast)_x = (\alpha_x)^\ast$. \end{prop} Proof - The proof is analogous to that of Proposition \ref{2cat5}. \hskip 3em \hbox{\BOX} \vskip 2ex Given 2-Hilbert spaces $H,H'$ and $L$, note that a bimorphism $T \colon H \times H' \to L$ induces a morphism \[ T^\ast \colon \hom(L,K) \to {\rm bihom}(H \times H',K). \] \begin{defn} \hspace{-0.08in}{\bf .}\hspace{0.1in} Given 2-Hilbert space $H,H'$, a {\rm tensor product} of $H$ and $H'$ is a bimorphism $T \colon H \times H' \to L$ together with a choice for each 2-Hilbert space $K$ of an equivalence of 2-Hilbert spaces extending $T^\ast \colon \hom(L,K) \to {\rm bihom}(H \times H',K)$. \end{defn} \noindent In the above situation, by abuse of language we may say simply that $T \colon H \times H' \to L$ is a tensor product of $H$ and $H'$. \begin{prop} \label{tensor} \hspace{-0.08in}{\bf .}\hspace{0.1in} Given finite-dimensional 2-Hilbert spaces $H$ and $H'$, there exists a tensor product $T \colon H \times H' \to L$. Given another tensor product $T' \colon H \times H' \to L'$, there is an equivalence $F \colon L \to L'$ for which the following diagram commutes up to a specified natural isomorphism: \[ \begin{diagram}[H \times H'] \node[2]{H \times H'} \arrow{sw,t}{T} \arrow{se,t}{T'} \\ \node{L} \arrow[2]{e,b}{F} \node[2]{L'} \end{diagram} \] \end{prop} Proof - Let $\{e_\lambda\}$ be a basis for $H$, and $\{f_\mu\}$ a basis for $H'$. Let $L$ be the skeletal 2-Hilbert space with a basis of objects denoted by $\{e_\lambda \otimes f_\mu\}$, and with \[ \hom(e_\lambda \otimes f_\mu,e_\lambda \otimes f_\mu) = \hom(e_\lambda,e_\lambda) \otimes \hom(f_\mu,f_\mu) \] as H*-algebras (using the obvious tensor product of H*-algebras). There is a unique bimorphism $T \colon H \times H' \to L$ with $T(e_\lambda,f_\mu) = e_\lambda \otimes f_\mu$. Given a 2-Hilbert space $K$ one may check that $T^\ast \colon \hom(L,K) \to {\rm bihom}(H \times H',K)$ extends to an equivalence. Choosing such an equivalence for every $K$ we obtain a tensor product of $H$ and $H'$. Given two tensor products as in the statement of the proposition, let $F\colon L \to L'$ be the image of $T'$ under the chosen equivalence ${\rm bihom}(H \times H', L') \simeq \hom(L,L')$. One can check that $L$ is an equivalence and that the above diagram commutes up to a specified natural isomorphism, much as in the usual proof that the tensor product of vector spaces is unique up to a specified isomorphism. \hskip 3em \hbox{\BOX} \vskip 2ex Given a tensor product of the 2-Hilbert spaces $H$ and $H'$, we often write its underlying 2-Hilbert space as $H \otimes H'$. This notation may tempt one to speak of `the' tensor product of $H$ and $H'$, which is is legitimate if one uses the generalized `the' as advocated by Dolan \cite{Dolan}. In a set, when we speak of `the' element with a given property, we implicitly mean that this element is unique. In a category, when we speak of `the' object with a given property, we merely mean that this object is unique up to isomorphism --- typically a specified isomorphism. Similarly, in a 2-category, when we speak of `the' object with a given property, we mean that this object is unique up to equivalence --- typically an equivalence that is specified up to a specified isomorphism. This is the sense in which we may refer to `the' tensor product of $H$ and $H'$. The generalized `the' may be extended in an obvious recursive fashion to $n$-categories. Suppose that $H$ and $H'$ are finite-dimensional 2-Hilbert spaces. Then for any pair of objects $x \in H$, $x' \in H'$, we can use the bimorphism $T \colon H \times H' \to H \otimes H'$ to define an object $x \otimes x' = T(x,x')$ in $H \otimes H'$. Similarly, given a morphism $f \colon x \to y$ in $H$ and a morphism $f' \colon x' \to y'$, we obtain a morphism \[ f \otimes f'\, \colon x \otimes x' \to y \otimes y'\] in $H \otimes H'$. We usually write \[ f \otimes x' \, \colon x \otimes x' \to y \otimes x' \] for the morphism $f \otimes 1_{x'}$, and \[ x \otimes f' \, \colon x \otimes x' \to x \otimes y' \] for the morphism $1_x \otimes f'$. We expect that $2{\rm Hilb}$ has the structure of a monoidal 2-category with the above tensor product as part of the monoidal structure. Kapranov and Voevodsky \cite{KV} have defined the notion of a weak monoidal structure on a strict 2-category, which should be sufficient for the purpose at hand. On the other hand the work of Gordon, Power and Street \cite{GPS} gives a fully general notion of weak monoidal 2-category, namely a 1-object tricategory. This should also be suitable for studying the tensor product on $2{\rm Hilb}$, though it might be considered overkill. Both these sorts of monoidal 2-category involve various extra structures besides the tensor product of objects in $2{\rm Hilb}$. Most of these should arise from the universal property of the tensor product. For example, suppose we are given a morphism $F \colon H \to H'$ and an object $K$ in $2{\rm Hilb}$. Thus we have bimorphisms $T \colon H \times K \to H \otimes K$ and $T' \colon H \times K' \to H \otimes K'$, and $T^\ast$ has some morphism \[ S \colon {\rm bihom}(H \times K, H' \otimes K) \to \hom(H \otimes K,H' \otimes K) \] as inverse up to natural isomorphism. Applying $S$ to the bimorphism given by the composite \[ \begin{diagram} [H \otimes K] \node{H \times K} \arrow{e,t}{F \times 1_K} \node{H' \times K} \arrow{e,t}{T'} \node{H' \otimes K} \end{diagram} \] we obtain a morphism we denote by \[ F \otimes K \colon H \otimes K \to H' \otimes K .\] Similarly, given an object $H \in 2{\rm Hilb}$ and a morphism $G \colon K \to K'$, we obtain a morphism \[ H \otimes G \colon H \otimes K \to H \otimes K'. \] Moreover, we have: \begin{prop} \label{tensorator} \hspace{-0.08in}{\bf .}\hspace{0.1in} Let $F \colon H \to H'$ and $G \colon K \to K'$ be morphisms in $2{\rm Hilb}$. Then the following diagram \[ \begin{diagram} \node{H \otimes K} \arrow{e,t}{F \otimes K} \arrow{s,l}{H \otimes G} \node{H' \otimes K} \arrow{s,r}{H' \otimes G} \\ \node{H \otimes K'} \arrow{e,b}{F \otimes K'} \node{H' \otimes K'} \end{diagram} \] commutes up to a specified natural isomorphism \[ {\bigotimes}_{F,G} \colon (F \otimes K)(H' \otimes G) \Rightarrow (H \otimes G)(F \otimes K') . \] \end{prop} Proof - Here we have fixed tensor products of all the 2-Hilbert spaces involved, so we have bimorphisms \[ T_{H,K} \colon H \times K \to H \otimes K \] and so on. Applying the equivalence \[ {\rm bihom}(H \times K, H' \otimes K') \simeq \hom(H \otimes K,H' \otimes K') \] coming from the definition of tensor product to the bimorphism given by the composite \begin{equation} \begin{diagram} [H \otimes K] \node{H \times K} \arrow{e,t}{F \times G} \node{H' \times K'} \arrow{e,t}{T_{H',K'}} \node{H' \otimes K'} \end{diagram} \label{bigotimes1} \end{equation} we obtain a morphism we denote by \[ F \otimes G \colon H \otimes K \to H' \otimes K'. \] We shall construct a natural isomorphism from $(F \otimes K)(H' \otimes G)$ to $F \otimes G$. Composing this with an analogous natural isomorphism from $F \otimes G$ to $(H \otimes G)(F \otimes K')$ one obtains $\bigotimes_{F,G}$. If we precompose $F \otimes G$ with $T_{H,K}$ we obtain a bimorphism naturally isomorphic to (\ref{bigotimes1}). If we precompose $(F \otimes K')(H \otimes G)$ with $T_{H,K}$, we obtain a bimorphism naturally isomorphic to \begin{equation} \begin{diagram} [H' \otimes K'.] \node{H \times K} \arrow{e,t}{F \times K} \node{H' \times K} \arrow{e,t}{T_{H',K}} \node{H' \otimes K} \arrow{e,t}{H' \otimes G} \node{H' \otimes K'} \end{diagram} \label{bigotimes2} \end{equation} Note also that in both cases, a {\it specified} natural isomorphism is given by the definition of tensor product. Since precomposition with $T_{H,K}$ is an equivalence between ${\rm bihom}(H \times K,H' \otimes K')$ and $\hom(H \otimes K,H' \otimes K'),$ it thus suffices to exhibit a natural isomorphism between (\ref{bigotimes1}) and (\ref{bigotimes2}). Factoring these by $F \times K$, it suffices to exhibit a natural isomorphism between \[ \begin{diagram} [H' \otimes K] \node{H' \times K} \arrow{e,t}{H' \times G} \node{H' \otimes K'} \arrow{e,t}{T_{H',K'}} \node{H' \otimes K'} \end{diagram} \] and \[ \begin{diagram} [H' \otimes K'.] \node{H' \times K} \arrow{e,t}{T_{H',K}} \node{H' \otimes K} \arrow{e,t}{H' \otimes G} \node{H' \otimes K'} \end{diagram} \] This arises from the definition of $H' \otimes G$. \hskip 3em \hbox{\BOX} \vskip 2ex \noindent The 2-morphism $\bigotimes_{F,G}$ is part of the structure one expects in a monoidal 2-category, and the fact that the diagram in Proposition \ref{tensorator} does not commute `on the nose' is one of the key ways in which monoidal 2-categories differ from monoidal categories. We expect a 2-categorical version of $\hom$-tensor adjointness to hold for the tensor product defined in this section and the $\hom$ defined in section \ref{dualityforobjects}. In other words, given finite-dimensional 2-Hilbert space $H, H',$ and $K$, the obvious functor from $\hom(H, \hom(H',K))$ to $\hom(H \otimes H', K)$ should be an equivalence. However, we shall not prove this here. \subsection{The braiding} The symmetry in ${\rm Cat}$ gives braiding morphisms in $2{\rm Hilb}$ as follows. Let $H$ and $H'$ be 2-Hilbert spaces. We may take their tensor product in either order, obtaining tensor products $T \colon H \times H' \to H \otimes H'$ and $T' \colon H' \times H \to H' \otimes H$. By the universal property of the tensor product, the bimorphism given by the composite \[ \begin{diagram} [H \otimes K] \node{H \times H'} \arrow{e,t}{S_{H,H'}} \node{H' \times H} \arrow{e,t}{T'} \node{H' \otimes H} \end{diagram} \] defines a morphism, the {\it braiding} \[ R_{H,H'} \colon H \otimes H' \to H' \otimes H .\] One can check that $R_{H,H'}$ is an equivalence. We expect that $2{\rm Hilb}$ has the structure of a braided monoidal 2-category with the above braiding morphisms. However, the existing notion of semistrict braided monoidal 2-category introduced by Kapranov and Voevodsky \cite{KV} and subsequently refined in HDA1 is insufficiently general to cover this example, since $2{\rm Hilb}$ is not a semistrict monoidal 2-category. One should however be able to strictify $2{\rm Hilb}$, obtaining a semistrict braided monoidal 2-category. Alternatively, the work of Trimble \cite{Trimble} should give a fully general notion of weak braided monoidal 2-category, namely a tetracategory with one object and one morphism. This should apply to $2{\rm Hilb}$ without strictification. In any event, both semistrict and weak braided monoidal 2-categories involve various structures in addition to the braiding morphisms. Most of these should arise from the universal property of the tensor product together with the properties of the symmetry in ${\rm Cat}$. For example, we have: \begin{prop} \label{braiding.naturalizer} \hspace{-0.08in}{\bf .}\hspace{0.1in} Let $F \colon H \to H'$ be a morphism and let $K$ be an object in $2{\rm Hilb}$. Then the following diagram \[ \begin{diagram} \node{H \otimes K} \arrow{e,t}{F \otimes K} \arrow{s,l}{R_{H,K}} \node{H' \otimes K} \arrow{s,r}{R_{H',K}} \\ \node{K \otimes H} \arrow{e,b}{K \otimes F} \node{K \otimes H'} \end{diagram} \] commutes up to a specified natural isomorphism \[ R_{F,K} \colon (F \otimes K) R_{H',K} \Rightarrow R_{H,K} (K \otimes F) .\] Similarly, given an object $H$ and a morphism $G \colon K \to K'$ in $2{\rm Hilb}$, the following diagram \[ \begin{diagram} \node{H \otimes K} \arrow{e,t}{H \otimes G} \arrow{s,l}{R_{H,K}} \node{H \otimes K'} \arrow{s,r}{R_{H,K'}} \\ \node{K \otimes H} \arrow{e,b}{G \otimes H} \node{K' \otimes H} \end{diagram} \] commutes up to a specified natural isomorphism \[ R_{H,G} \colon (H \otimes G)R_{H,K'} \Rightarrow R_{H,K}(G \otimes H) .\] \end{prop} Proof - We only treat the first case as the second is analogous. Applying the equivalence \[ {\rm bihom}(H \times K, K' \otimes H') \simeq \hom(H \otimes K,K' \otimes H') \] coming from the definition of tensor product to the bimorphism given by the composite \begin{equation} \begin{diagram} [H \otimes K] \node{H \times K} \arrow{e,t}{F \times K} \node{H' \times K} \arrow{e,t}{S_{H',K}} \node{K \times H'} \arrow{e,t}{T_{K,H'}} \node{K \otimes H'} \end{diagram} \label{braiding.nat.1} \end{equation} we obtain a morphism we denote by $A \colon H \otimes K \to K' \otimes H$. We shall construct a natural isomorphism from $(F \otimes K)R_{H',K}$ to $A$. Using the fact that (\ref{braiding.nat.1}) equals \[ \begin{diagram} [H \otimes K] \node{H \times K} \arrow{e,t}{S_{H,K}} \node{K \times H} \arrow{e,t}{K \times F} \node{K \times H'} \arrow{e,t}{T_{K,H'}} \node{K \otimes H'} \end{diagram} \] one can similarly obtain a natural isomorphism from $A$ to $R_{H,K}(K \otimes F)$. The composite of these is $R_{F,K}$. If we precompose $A$ with $T_{H,K}$ we obtain a bimorphism naturally isomorphic to (\ref{braiding.nat.1}). If we precompose $(F \otimes K)R_{H',K}$ with $T_{H,K}$, we obtain a bimorphism naturally isomorphic to \begin{equation} \begin{diagram} [H' \otimes K.] \node{H \times K} \arrow{e,t}{F \times K} \node{H' \times K} \arrow{e,t}{T_{H',K}} \node{H' \otimes K} \arrow{e,t}{R_{H',K}} \node{K' \otimes H} \end{diagram} \label{braiding.nat.2} \end{equation} In both cases, a natural isomorphism is given by the definition of tensor product. It thus suffices to exhibit a natural isomorphism between (\ref{braiding.nat.1}) and (\ref{braiding.nat.2}). This may be constructed as in the proof of Proposition \ref{tensorator}. \hskip 3em \hbox{\BOX} \vskip 2ex \subsection{The involutor} As indicated in Figure 1, for $2{\rm Hilb}$ to be a stable 2-category it should possess an extra layer of structure after the tensor product and the braiding, namely the `involutor'. Also, this structure should have an extra property making $2{\rm Hilb}$ `strongly involutory'. The involutor is a weakened form of the equation appearing in the definition of a symmetric monoidal category. Namely, while the braiding need not satisfy \[ R_{H',H} R_{H,H'} = 1_{H \otimes H'} \] for all objects $H,H' \in 2{\rm Hilb}$, there should be a 2-isomorphism \[ I_{H,H'} \colon R_{H,H'} R_{H',H} \Rightarrow 1_{H \otimes H'}, \] the involutor. We construct the involutor as follows. Choose tensor products $T \colon H \times H' \to H \otimes H'$ and $T' \colon H' \times H \to H' \otimes H$. Then by the universality of the tensor product, the commutativity of \[ \begin{diagram} [H \times H'] \node[2]{H' \times H} \arrow{se,t}{S_{H',H}} \\ \node{H \times H'} \arrow[2]{e,b}{1_{H \times H'}} \arrow{ne,t}{S_{H,H'}} \node[2]{H \times H'} \\ \end{diagram} \] implies that \[ \begin{diagram} [H \times H'] \node[2]{H' \otimes H} \arrow{se,t}{R_{H',H}} \\ \node{H \otimes H'} \arrow[2]{e,b}{1_{H \otimes H'}} \arrow{ne,t}{R_{H,H'}} \node[2]{H \otimes H'} \\ \end{diagram} \] commutes up to a specified natural transformation. This is the involutor \[ I_{H,H'} \colon R_{H,H'} R_{H',H} \Rightarrow 1_{H \otimes H'}. \] In addition, for $2{\rm Hilb}$ to be stable, or `strongly involutory', the involutor should satisfy a special coherence law of its own, in analogy to how the braiding satisfies a special equation in a symmetric monoidal category. In HDA0 this equation was described in terms of $R_{H,H'}$ and a weak inverse thereof, but it turns out to be easier to give the equation by stating that the following horizontal composites agree: \[ I_{H,H'} \circ 1_{R_{H,H'}} \colon R_{H,H'} R_{H',H} R_{H,H'} \Rightarrow R_{H,H'} \] and \[ 1_{R_{H,H'}} \circ I_{H,H'} \colon R_{H,H'} R_{H',H} R_{H,H'} \Rightarrow R_{H,H'} \] This is indeed the case, as one can show using the properties of the tensor product. \section{2-H*-algebras}\label{2-H*-algebras} Now we consider 2-Hilbert spaces with extra structure and properties, as listed in the second column of Figure 2. \begin{defn} \hspace{-0.08in}{\bf .}\hspace{0.1in} A {\rm 2-H*-algebra} $H$ is a 2-Hilbert space equipped with a {\rm product} bimorphism $\otimes \colon H \times H \to H$, a {\rm unit} object $1 \in H$, a unitary natural transformation $a_{x,y,z} \colon (x \otimes y) \otimes z \to x \otimes (y \otimes z)$ called the {\rm associator}, and unitary natural transformations $\ell_x \colon 1 \otimes x \to x$, $r_x \colon x \otimes 1 \to x$ called the {\rm left} and {\rm right unit laws}, making $H$ into a monoidal category. We require also that every object $x \in H$ has a left dual. \end{defn} \noindent Recall that for $H$ to be a monoidal category, one demands that the following pentagon commute: \[ \begin{diagram}[((x \otimes y)\otimes z)\otimes w] \node{((x \otimes y) \otimes z)\otimes w} \arrow{e,t}{a_{x\otimes y,z,w}} \arrow{s,l}{a_{x,y,z}\otimes w} \node{(x\otimes y)\otimes(z\otimes w)} \arrow{e,t}{a_{x,y,z\otimes w}} \node{x \otimes (y\otimes(z \otimes w))} \\ \node{(x \otimes (y \otimes z)) \otimes w} \arrow[2]{e,t}{a_{x,y\otimes z,w}} \node[2]{x \otimes ((y \otimes z)\otimes w)} \arrow{n,r}{x \otimes a_{y,z,w}} \end{diagram} \] as well as the following diagram involving the unit laws: \[ \begin{diagram}[(1 \otimes x) \otimes 1] \node{(1 \otimes x) \otimes 1} \arrow{s,l}{\ell_x \otimes 1} \arrow[2]{e,t}{a_{1,x,1}} \node[2]{1 \otimes (x\otimes 1)} \arrow{s,r}{1 \otimes r_x} \\ \node{x \otimes 1} \arrow{e,t}{r_x} \node{x} \node{1 \otimes x} \arrow{w,t}{\ell_x} \end{diagram} \] Mac Lane's coherence theorem \cite{MacLane2} says that every monoidal category is equivalent, as a monoidal category, to a {\it strict} monoidal category, that is, one for which the associators and unit laws are all identity morphisms. Sometimes we will use this to streamline formulas by not parenthesizing tensor products and not writing the associators and unit laws. Such formulas apply literally only to the strict case, but one can always use Mac Lane's theorem to apply them to general monoidal categories. In practice, this amounts to parenthesizing tensor products however one likes, and inserting associators and unit laws when needed to make the formulas make sense. A {\it left dual} of an object $x$ in a monoidal category is an object $y$ together with morphisms \[ e \colon y \otimes x \to 1 \] and \[ i \colon 1 \to x \otimes y ,\] called the {\it unit} and {\it counit}, such that the following diagrams commute: \[ \begin{diagram} [x \otimes y \otimes x] \node{x} \arrow[2]{e,t}{1_x} \arrow{se,b}{i \otimes x} \node[2]{x} \\ \node[2]{x \otimes y \otimes x} \arrow{ne,b}{x \otimes e} \end{diagram} \] \[ \begin{diagram} [x \otimes y \otimes x] \node{y} \arrow[2]{e,t}{1_y} \arrow{se,b}{y \otimes i} \node[2]{y} \\ \node[2]{y \otimes x \otimes y} \arrow{ne,b}{e \otimes y} \end{diagram} \] (These diagrams apply literally only when the monoidal category is strict.) In this situation we also say that $x$ is a {\it right dual} of $y$, and that $(x,y,i,e)$ is an {\it adjunction}. All adjunctions having $x$ as right dual are uniquely isomorphic in the following sense: \begin{prop} \label{2H*1} \hspace{-0.08in}{\bf .}\hspace{0.1in} Given an adjunction $(x,y,i,e)$ in a monoidal category and an isomorphism $f\colon y \to y'$, there is an adjunction $(x,y',i',e')$ given by: \[ i' = i(x \otimes f), \qquad e' = (f^{-1} \otimes x)e.\] Conversely, given two adjunctions $(x,y,i,e)$ and $(x,y',i',e')$, there is a unique isomorphism $f \colon y \to y'$ for which $i' = i(x \otimes f)$ and $e' = (f^{-1} \otimes x)e$. This is given in the strict case by the composite \[ \begin{diagram}[y \otimes (x \otimes y')] \node{y = y \otimes 1} \arrow{e,t}{y \otimes i'} \node{y \otimes x \otimes y'} \arrow{e,t}{e \otimes y'} \node{1 \otimes y' = y'} \end{diagram} \] \end{prop} Proof - This result is well-known and the proof is a simple calculation. \hskip 3em \hbox{\BOX} \vskip 2ex \noindent Similarly, any two adjunctions having a given object as right dual are canonically isomorphic. We may thus speak of `the' left or right dual of a given object, using the generalized `the', as described in Section \ref{tensorproduct}. Note that duality at the morphism level of a 2-H*-algebra allows us to turn left duals into right duals, and vice versa, at the object level: \begin{prop}\label{2H*2}\hspace{-0.08in}{\bf .}\hspace{0.1in} Suppose that $H$ is a 2-H*-algebra. Then $(x,x^\ast,i,e)$ is an adjunction if and only if $(x^\ast,x,e^\ast,i^\ast)$ is an adjunction. \end{prop} Proof - The proof is analogous to that of Proposition \ref{2cat1}. \hskip 3em \hbox{\BOX} \vskip 2ex Next we turn to braided and symmetric 2-H*-algebras. A good example of a braided 2-H*-algebra is the category of tilting modules of a quantum group when the parameter $q$ is a suitable root of unity \cite{CP}. Categories very similar to our braided 2-H*-algebras have been studied by Fr\"ohlich and Kerler \cite{FK} under the name `C*-quantum categories'; our definitions differ only in some fine points. A good example of a symmetric 2-H*-algebra is the category of finite-dimensional continuous unitary representations of a compact topological group. Doplicher and Roberts \cite{DR} have studied categories very similar to our symmetric 2-H*-algebras. \begin{defn} \hspace{-0.08in}{\bf .}\hspace{0.1in} A {\rm braided 2-H*-algebra} is a 2-H*-algebra $H$ equipped with a unitary natural isomorphism $B_{x,y} \colon x \otimes y \to y \otimes x$ making $H$ into a braided monoidal category. \end{defn} \begin{defn} \hspace{-0.08in}{\bf .}\hspace{0.1in} A {\rm symmetric 2-H*-algebra} is a 2-H*-algebra for which the braiding is a symmetry. \end{defn} Recall that for $H$ to be a braided monoidal category, the following two hexagons must commute: \[ \begin{diagram}[(x \otimes y) \otimes z)] \node{x \otimes (y \otimes z)} \arrow{s,l}{B_{x,y \otimes z}} \arrow{e,t}{a^{-1}_{x,y,z}} \node{(x \otimes y) \otimes z} \arrow{e,t}{B_{x,y} \otimes z} \node{(y \otimes x) \otimes z} \arrow{s,r}{a_{y,x,z}} \\ \node{(y \otimes z) \otimes x} \arrow{e,b}{a_{y,z,x}} \node{y \otimes (z \otimes x)} \arrow{e,t}{y \otimes B_{x,z}} \node{y \otimes (x \otimes z)} \end{diagram} \] \[ \begin{diagram}[(x \otimes y) \otimes z)] \node{(x \otimes y) \otimes z} \arrow{s,l}{B_{x \otimes y, z}} \arrow{r,t}{a_{x,y,z}} \node{x \otimes (y \otimes z)} \arrow{e,t}{x \otimes B_{y,z}} \node{x \otimes (z \otimes y)} \arrow{s,r}{a_{x,z,y}^{-1}} \\ \node{z \otimes (x \otimes y)} \arrow{e,b}{a^{-1}_{z,x,y}} \node{(z \otimes x) \otimes y} \arrow{e,t}{B_{x,z} \otimes y} \node{(x \otimes z) \otimes y} \end{diagram} \] as well as the following diagrams: \[ \begin{diagram}[1 \otimes x] \node{1 \otimes x} \arrow{se,b}{\ell_x} \arrow[2]{e,t}{B_{1,x}} \node[2]{x \otimes 1} \arrow{sw,b}{r_x} \\ \node[2]{x} \end{diagram} \] \[ \begin{diagram}[1 \otimes x] \node{x \otimes 1} \arrow{se,b}{r_x} \arrow[2]{e,t}{B_{x,1}} \node[2]{1 \otimes x} \arrow{sw,b}{\ell_x} \\ \node[2]{x} \end{diagram} \] The braiding is a symmetry if $B_{x,y} = B_{y,x}^{-1}$ for all objects $x$ and $y$. \subsection{The balancing} \label{balancing} In the study of braided monoidal categories where objects have duals, it is common to introduce something called the `balancing'. The balancing can treated in various ways \cite{FK,JS2,RT}. For example, one may think of it as a choice of automorphism $b_x \colon x \to x$ for each object $x$, which is required to satisfy certain laws. While very important in topology, this extra structure seems somewhat ad hoc and mysterious from the algebraic point of view. We now show that braided 2-H*-algebras are automatically equipped with a balancing. The reason is that not only the objects, but also the morphisms, have duals. In fact, some of what follows would apply to any braided monoidal category in which both objects and morphisms have duals. In any 2-H*-algebra, Proposition \ref{2H*2} gives a way to make any object $x$ into the left dual of its left dual $x^\ast$. In a braided 2-H*-algebra, $x$ also becomes the left dual of $x^\ast$ in another way: \begin{prop} \label{2H*2.5} \hspace{-0.08in}{\bf .}\hspace{0.1in} Let $H$ be a braided 2-H*-algebra. Then $(x,x^\ast,i,e)$ is an adjunction if and only if $(x^\ast,x,iB_{x,x^\ast},B_{x,x^\ast}e)$ is an adjunction. \end{prop} Proof - The proof is a simple computation. \hskip 3em \hbox{\BOX} \vskip 2ex It follows from Proposition \ref{2H*1} that these two ways to make $x$ into the left dual of $x^\ast$ determine an automorphism of $x$. Simplifying the formula for this automorphism somewhat, we make the following definition: \begin{defn}\hspace{-0.08in}{\bf .}\hspace{0.1in} If $H$ is a braided 2-H*-algebra and $(x,x^\ast,i,e)$ is an adjunction in $H$, the {\rm balancing} of the adjunction is the morphism $b \colon x \to x$ given in the strict case by the composite: \[ \begin{diagram} [(x \otimes x^\ast) \otimes x] \node{x} \arrow{e,t}{e^\ast \otimes x} \node{x^\ast \otimes x \otimes x} \arrow{e,t}{x^\ast \otimes B_{x,x}} \node{x^\ast \otimes x \otimes x} \arrow{e,t}{e \otimes x} \node{x} \end{diagram} \] \end{defn} It is perhaps easiest to understand the significance of the balancing in terms of its relation to topology. We shall be quite sketchy about describing this, but the reader can fill in the details using the ideas described in HDA0 and the many references therein. Especially relevant is the work of Freyd and Yetter \cite{FY}, Joyal and Street \cite{JS}, and Reshetikhin and Turaev \cite{RT,T}. We discuss this relationship more carefully in the Conclusions. \bigskip \centerline{\epsfysize=1.5in\epsfbox{2dtangle.eps}} \medskip \centerline{3. Typical tangle in 2 dimensions} \medskip The basic idea is to use tangles to represent certain morphisms in 2-H*-algebras. A typical oriented tangle in 2 dimensions is shown in Figure 3. If we fix an adjunction $(x,x^\ast,i,e)$ in a strict 2-H*-algebra $H$, any such tangle corresponds uniquely to a morphism in $H$ as follows. As shown in Figure 4, vertical juxtaposition of tangles corresponds to the composition of morphisms, while horizontal juxtaposition corresponds to the tensor product of morphisms. \vbox{ \bigskip \centerline{\epsfysize=1.5in\epsfbox{products.eps}} \medskip \centerline{4. Composition and tensor product of tangles} \medskip } \noindent Thus it suffices to specify the morphisms in $H$ corresponding to certain basic tangles from which all others can be built up by composition and tensor product. These basic tangles are shown in Figures 5 and 6. A downwards-pointing segment corresponds to the identity on $x$, while an upwards-pointing segment corresponds to the identity on $x^\ast$. \bigskip \centerline{\epsfysize=0.75in\epsfbox{identities.eps}} \medskip \centerline{5. Tangles corresponding to $1_x$ and $1_{x^\ast}$} \medskip \noindent The two oriented forms of a `cup' tangle correspond to the morphisms $e$ and $i^\ast$, while the two oriented forms of a `cap' correspond to $i$ and $e^\ast$. \bigskip \centerline{\epsfysize=0.75in\epsfbox{unitcounit.eps}} \medskip \centerline{6. Tangles corresponding to $e$, $i^\ast$, $i$, and $e^\ast$} \medskip \noindent It then turns out that isotopic tangles correspond to the same morphism in $H$. The main thing to check is that the isotopic tangles shown in Figure 7 correspond to the same morphisms. This follows from the triangle diagrams in the definition of an adjunction. Similar equations with the orientation of the arrows reversed follow from Proposition \ref{2H*2}. \bigskip \centerline{\epsfysize=1.5in\epsfbox{triangle.eps}} \medskip \centerline{7. Tangle equations corresponding to the definition of adjunction} \medskip If $H$ is braided, we can also map framed oriented tangles in 3 dimensions to morphisms in $H$. A typical such tangle is shown in Figure 8. We use the blackboard framing, in which each strand is implicitly equipped with a vector field normal to the plane in which the tangle is drawn. \bigskip \centerline{\epsfysize=1.5in\epsfbox{3dtangle.eps}} \medskip \centerline{8. Typical tangle in 3 dimensions} \medskip \noindent We interpret the basic tangles in Figures 5 and 6 as we did before. Moreover, we let the tangles in Figure 9 correspond to the morphisms $B_{x,x}$, $B_{x^\ast,x}$, $B_{x,x^\ast}$, and $B_{x^\ast,x^\ast}$, and let the tangles in Figure 10 correspond to the morphisms $B_{x,x}^{-1}$, $B_{x^\ast,x}^{-1}$, $B_{x,x^\ast}^{-1}$, and $B_{x^\ast,x^\ast}^{-1}$. \bigskip \centerline{\epsfysize=1in\epsfbox{braid1.eps}} \medskip \centerline{7. Tangles corresponding to $B_{x,x}$, $B_{x^\ast,x}$, $B_{x,x^\ast}$, and $B_{x^\ast,x^\ast}$} \medskip \bigskip \centerline{\epsfysize=1in\epsfbox{braid2.eps}} \medskip \centerline{8. Tangles corresponding to $B_{x,x}^{-1}$, $B_{x^\ast,x}^{-1}$, $B_{x,x^\ast}^{-1}$, and $B_{x^\ast,x^\ast}^{-1}$} \medskip Now suppose we wish isotopic framed oriented tangles to correspond to the same morphism in $H$. Invariance under the 2nd and 3rd Reidemeister moves follows from the properties of the braiding, so it suffices to check invariance under the framed version of the 1st Reidemeister move. For this, note that the tangle shown in Figure 9 corresponds to the balancing of the adjunction $(x,x^\ast,i,e)$. This tangle has a $2\pi$ twist in its framing. \vbox{ \bigskip \centerline{\epsfysize=1.5in\epsfbox{balancing.eps}} \medskip \centerline{9. Tangle corresponding to the balancing $b \colon x \to x$} \medskip } \noindent The framed version of the 1st Reidemeister move, shown in Figure 10, represents the cancellation of two opposite $2\pi$ twists in the framing. Both tangles in this picture correspond to the same morphism in $H$ precisely when the balancing $b \colon x \to x$ is unitary. \bigskip \centerline{\epsfysize=1.5in\epsfbox{unitarity.eps}} \medskip \centerline{10. Tangle equation corresponding to unitarity of the balancing} \medskip In short, we obtain a map from isotopy classes of framed oriented tangles in 3 dimensions to morphisms in a braided 2-H*-algebra $H$ whenever we choose an adjunction in $H$ whose balancing is unitary. This motivates the following definition: \begin{defn} \hspace{-0.08in}{\bf .}\hspace{0.1in} An adjunction $(x,x^\ast,i,e)$ in a braided 2-H*-algebra is {\rm well-balanced} if its balancing is unitary. \end{defn} \noindent Similarly, given any well-balanced adjunction in a symmetric 2-H*-algebra $H$, we obtain a map from isotopy classes of framed oriented tangles in 4 dimensions to morphisms in $H$. We may draw tangles in 4 dimensions just as we draw tangles in 3 dimensions, but there is an extra rule saying that any right-handed crossing is isotopic to the corresponding left-handed crossing. One case of this rule is shown in Figure 11. Invariance under these isotopies follows directly from the fact that the braiding is a symmetry. \bigskip \centerline{\epsfysize=1.0in\epsfbox{symmetry.eps}} \medskip \centerline{11. Tangle equation corresponding to symmetry} \medskip The important fact is that well-balanced adjunctions exist and are unique up to a unique unitary isomorphism. Moreover, all of them have the same balancing: \begin{thm}\label{2H*3} \hspace{-0.08in}{\bf .}\hspace{0.1in} Suppose $H$ is a braided 2-H*-algebra. For every object $x \in H$ there exists a well-balanced adjunction $(x,y,i,e)$. Given well-balanced adjunctions $(x,y,i,e)$ and $(x,y',i',e')$, there is a unique morphism $u \colon y \to y'$ such that \[ i' = i(u \otimes x), \qquad e' = (x \otimes u^{-1})e, \] and this morphism is unitary. \end{thm} Proof - To simplify notation we assume without loss of generality that $H$ is strict. Suppose first that $x \in H$ is simple. Then for any adjunction $(x,y,i,e)$, the balancing equals $\beta 1_x$ for some nonzero $\beta \in {\Bbb C}$. By Proposition \ref{2H*1} we may define a new adjunction $(x,y,|\beta|^{1/2}i,|\beta|^{-1/2}e)$. Since the balancing of this adjunction equals $\beta |\beta|^{-1} 1_x$, this adjunction is well-balanced. Next suppose that $x \in H$ is arbitrary. Using Theorem \ref{H*3} we can write $x$ as an orthogonal direct sum of simple objects $x_j$, in the sense that there are morphisms \[ p_j \colon x \to x_j \] with \[ p_j^\ast p_j = 1_{x_j}, \qquad \sum_j p_j p_j^\ast = 1_x. \] Let $y_j$ be a left dual of $x_j$, and define $y$ to be an orthogonal direct sum of the objects $y_j$, with morphisms \[ q_j \colon y \to y_j \] such that \[ q_j^\ast q_j = 1_{y_j}, \qquad \sum_j q_j q_j^\ast = 1_y . \] Since the $x_j$ are simple, there exist adjunctions $(x_j,y_j,e_j,i_j)$ for which the balancings $b_j \colon x_j \to x_j$ are unitary. Define the adjunction $(x,y,i,e)$ by \[ i = \sum_j i_j (p_j^\ast \otimes q_j^\ast) , \qquad e = \sum_j (q_j \otimes p_j) e_j .\] One can check that this is indeed an adjunction and that the balancing $b \colon x \to x$ of this adjunction is given by \[ b = \sum_j p_j b_j p_j^\ast ,\] and is therefore unitary. Now suppose that $(x,y',i',e')$ is any other well-balanced adjunction with $x$ as right dual. Let $b'$ denote the balancing of this adjunction. We shall prove that $b' = b$. By Propositions \ref{H*4} and \ref{2H*1} there exists a unitary morphism $g \colon y \to y'$, and we have \begin{eqnarray*} b' &=& (e'^\ast \otimes x)(y' \otimes B_{x,x})(e' \otimes x) \\ &=& (e'^\ast(g^\ast \otimes x) \otimes x)(y \otimes B_{x,x}) ((g \otimes x)e' \otimes x). \end{eqnarray*} By Proposition \ref{2H*1}, $(x,y,(g \otimes x)e',i(x \otimes g^{-1}))$ is an adjunction, so by the uniqueness up to isomorphism of right adjoints we have $(g \otimes x)e' = (y \otimes f)e$ for some isomorphism $f \colon x \to x$. We thus have \begin{eqnarray*} b' &=& (e^\ast(y \otimes f^\ast) \otimes x)(y \otimes B_{x,x}) ((y \otimes f)e \otimes x) \\ &=& fbf^\ast .\end{eqnarray*} We may write $x$ as an orthogonal direct sum \[ x = \bigoplus_\lambda x_\lambda \] where $\{e_\lambda\}$ is a basis of $H$ and $x_\lambda$ is a direct sum of some number of copies of $e_\lambda$. Then by our previous formula for $b$ we have \[ b = \bigoplus_\lambda \beta_\lambda 1_{x_\lambda} \] with $|\beta_\lambda| = 1$ for all $\lambda$. We also have \[ f = \bigoplus_\lambda f_\lambda \] for some morphisms $f_\lambda \colon x_\lambda \to x_\lambda$. It follows that \[ b' = \bigoplus_\lambda \beta_\lambda f_\lambda {f_\lambda}^\ast \] Since $b$ and $b'$ are unitary it follows that each morphism $f_\lambda {f_\lambda}^\ast$ is unitary. Since the only positive unitary operator is the identity, using Theorem \ref{H*3} it follows that each $f_\lambda {f_\lambda}^\ast$ is the identity, so $b' = b$ as desired. By Proposition \ref{2H*1}, we know there is a unique isomorphism $u \colon y \to y'$ with \[ i' = i(u \otimes x) , \qquad e' = (x \otimes u^{-1})e ,\] and we need to show that $u$ is unitary. Since $b' = b$, we have \[ (ib' \otimes y)(x \otimes B_{y,y}^{-1})(i^\ast \otimes y)= (ib \otimes y)(x \otimes B_{y,y}^{-1})(i^\ast \otimes y) ,\] and if one simplifies this equation using the fact that \[ b = (e^\ast \otimes x)(y' \otimes B_{x,x})(e \otimes x) \] and \[ b' = (e^\ast(x \otimes (u^{-1})^\ast) \otimes x)(y' \otimes B_{x,x}) ((x \otimes u^{-1})e \otimes x),\] one finds that $u$ is unitary. \hskip 3em \hbox{\BOX} \vskip 2ex \begin{cor} \hspace{-0.08in}{\bf .}\hspace{0.1in} In a braided 2-H*-algebra every well-balanced adjunction with $x$ as right dual has the same balancing, which we call {\rm the balancing} of $x$ and denote as $b_x \colon x \to x$. \end{cor} Proof - This was shown in the proof above. \hskip 3em \hbox{\BOX} \vskip 2ex Note that for any simple object $x$ in a braided 2-H*-algebra, the balancing $b_x$ must equal $1_x$ times some unit complex number, the {\it balancing phase} of $x$. In physics, the balancing phase describes the change in the wavefunction of a particle that undergoes a $2\pi$ rotation. Note that in a symmetric 2-H*-algebra \begin{eqnarray*} b_x &=& (e_x^\ast \otimes 1_x)(1_{x^\ast} \otimes B_{x,x}) (e_x \otimes 1_x) \\ &=& (e_x^\ast \otimes 1_x)(1_{x^\ast} \otimes B^\ast_{x,x}) (e_x \otimes 1_x) \\ &=& b_x^\ast, \end{eqnarray*} so $b_x^2 = 1_x$. Thus in this case the balancing phase of any simple object must be $\pm 1$. In physics, this corresponds to the fact that particles in 4-dimensional spacetime are either bosons and fermions depending on the phase they acquire when rotated by $2\pi$, while in 3-dimensional spacetime other possibilities, sometimes called `anyons', can occur \cite{DR,FK}. More generally, we make the following definition: \begin{defn} \hspace{-0.08in}{\bf .}\hspace{0.1in} If $H$ is a symmetric 2-H*-algebra, an object $x \in H$ is {\rm even} or {\rm bosonic} if $b_x = 1$, and {\rm odd} or {\rm fermionic} if $b_x = -1$. We say $H$ is {\rm even} or {\rm purely bosonic} if every object $x \in H$ is even. \end{defn} \noindent Note that if $x \oplus y$ is an orthogonal direct sum, \[ b_{x \oplus y} = b_x \oplus b_y, \] so an object in any symmetric 2-H*-algebra is even (resp.\ odd) if and only if it is a direct sum of even (resp.\ odd) simple objects. Also, since \[ b_{x \otimes y} = (b_x \otimes b_y)B_{x,y}B_{y,x}, \] it follows that the tensor product of two even or two odd objects is even, while the tensor product of an even and an odd object is odd. There is a way to turn any symmetric 2-H*-algebra into an even one, which will be useful in Section \ref{recon}. \begin{prop} \label{bosonization} \hspace{-0.08in}{\bf .}\hspace{0.1in} {\rm (Doplicher-Roberts)}\, Suppose $H$ is a symmetric 2-H*-algebra. Then there is a braiding $B'$ on $H$ given on simple objects $x,y \in H$ by \[ B^\flat_{x,y} = (-1)^{|x|\,|y|} B_{x,y} \] where $|x|$ equals $0$ or $1$ depending on whether $x$ is even or odd, and similarly for $|y|$. Let $H^\flat$ denote $H$ equipped with the new braiding $B^\flat$. Then $H^\flat$ is an even symmetric 2-H*-algebra, the {\rm bosonization} of $H$. \end{prop} Proof - This is a series of straightforward computations. One approach involves noting that for any objects $x,y \in H$, \[ B^\flat_{x,y} = {1\over 2}\, B_{x,y} (1_x \otimes 1_y + 1_x \otimes b_y + b_x \otimes 1_y - b_x \otimes b_y) . \] \hbox{\hskip 30em} \hskip 3em \hbox{\BOX} \vskip 2ex \noindent The above proposition is essentially due to Doplicher and Roberts, who proved it in a slightly different context \cite{DR}. However, the term `bosonization' is borrowed from Majid \cite{Majid}, who uses it to denote a related process that turns a super-Hopf algebra into a Hopf algebra. \subsection{Trace and dimension} The notion of the `dimension' of an object in a braided 2-H*-algebra will be very important in Section \ref{recon}. First we introduce the related notion of `trace'. \begin{defn}\hspace{-0.08in}{\bf .}\hspace{0.1in} If $H$ is a braided 2-H*-algebra and $f \colon x \to x$ is a morphism in $H$, for any well-balanced adjunction $(x,x^\ast,i,e)$ we define the {\rm trace} of $f$, ${\rm tr}(f) \in {\rm end}(1)$, by \[ {\rm tr}(f) = e (x^\ast \otimes f) e^\ast .\] \end{defn} \noindent The trace is independent of the choice of well-balanced adjunction, by Theorem \ref{2H*3}. Also, one can show that an obvious alternative definition of the trace is actually equivalent: \[ {\rm tr}(f) = i^\ast (f \otimes x^\ast) i .\] \begin{defn}\hspace{-0.08in}{\bf .}\hspace{0.1in} If $H$ is a braided 2-H*-algebra, we define the {\it dimension} of $x$, $\dim(x)$, to be ${\rm tr}(1_x)$. \end{defn} \noindent Note that $x,y$ are objects in a braided 2-H*-algebra, we have \[ \dim(x \oplus y) = \dim(x) + \dim(y), \qquad \dim(x \otimes y) = \dim(x) \dim(y), \qquad \dim(x^\ast) = \dim(x) .\] Moreoever, we have: \begin{prop} \label{dim} \hspace{-0.08in}{\bf .}\hspace{0.1in} If $H$ is a symmetric 2-H*-algebra and $x \in H$ is any object, then the spectrum of $\dim(x)$ is a subset of ${\Bbb N} = \{0,1,2,\dots\}$. \end{prop} Proof - We follow the argument of Doplicher and Roberts \cite{DR}. For any $n \ge 0$, the group algebra of the symmetric group $S_n$ acts as endomorphisms of $x^{\otimes n}$, and the morphisms $p_S, p_A$ corresponding to complete symmetrization and complete antisymmetrization, respectively, are self-adjoint projections in the H*-algebra ${\rm end}(x^{\otimes n})$. It follows that ${\rm tr}(p_S), {\rm tr}(p_A) \ge 0$. If $x$ is even, a calculation shows that \[ {\rm tr}(p_A) = {1\over n!}\, \dim(x)(\dim(x) - 1)\cdots (\dim(x) - n + 1) \] For this to be nonnegative for all $n$, the spectrum of $\dim(x)$ must lie in ${\Bbb N}$. Similarly, if $x$ is odd, a calculation shows that \[ {\rm tr}(p_S) = {1\over n!} \, \dim(x)(\dim(x) - 1)\cdots (\dim(x) - n + 1) \] so again the spectrum of $\dim(x)$ lies in ${\Bbb N}$. In general, any object $\dim(x)$ is a sum of simple objects, which are either even or odd, so by the additivity of dimension, the spectrum of $\dim(x)$ again lies in ${\Bbb N}$. \hskip 3em \hbox{\BOX} \vskip 2ex For any 2-H*-algebra, the Eckmann-Hilton argument shows that ${\rm end}(1)$ is a commutative H*-algebra, and thus isomorphic to a direct sum of copies of ${\Bbb C}$. (See HDA0 or HDA1 for a explanation of the Eckmann-Hilton argument.) \begin{defn} \hspace{-0.08in}{\bf .}\hspace{0.1in} A 2-H*-algebra $H$ is {\rm connected} if the unit object $1 \in H$ is simple. \end{defn} \noindent In a connected 2-H*-algebra, ${\rm end}(1) \cong {\Bbb C}$. The dimension of any object in a connected symmetric 2-H*-algebra is thus a nonnegative integer. In addition to the above notion of dimension it is also interesting to consider the `quantum dimension'. Here our treatment most closely parallels that of Majid \cite{Majid}. \begin{defn} \hspace{-0.08in}{\bf .}\hspace{0.1in} If $H$ is a braided 2-H*-algebra and $f \colon x \to x$ is a morphism in $H$, for any well-balanced adjunction $(x,x^\ast,i,e)$ we define the {\rm quantum trace} of $f$, ${\rm qtr}(f) \in {\rm end}(1)$, by \[ {\rm qtr}(f) = {\rm tr}(b_x f) .\] We define the {\rm quantum dimension} of $x$, ${\rm qdim}(x)$, to be ${\rm qtr}(1_x)$. \end{defn} \noindent In the case of a symmetric 2-H*-algebra, the quantum trace is also called the `supertrace'. Suppose $H$ is a connected symmetric 2-H*-algebra and $x$ is a simple object. Then ${\rm qdim}(x) \ge 0$ if $x$ is even and ${\rm qdim}(x) \le 0$ if $x$ is odd. The idea of odd objects as negative-dimensional is implicit in Penrose's work on negative-dimensional vector spaces \cite{Penrose}. \subsection{Homomorphisms and 2-homomorphisms} There is a 2-category with 2-H*-algebras as objects and `homomorphisms' and `2-homomorphisms' as morphisms and 2-morphisms, respectively. This is also true for braided 2-H*-algebras and symmetric 2-H*-algebras. \begin{defn} \hspace{-0.08in}{\bf .}\hspace{0.1in} Given 2-H*-algebras $H$ and $H'$, a {\rm homomorphism} $F \colon H \to H'$ is a morphism of 2-Hilbert spaces that is also a monoidal functor. If $H$ and $H'$ are braided, we say that $F$ is a {\rm homomorphism of braided 2-H*-algebras} if $F$ is additionally a braided monoidal functor. If $H$ and $H'$ are symmetric, we say that $F$ is a {\rm homomorphism of symmetric 2-H*-algebras} if $F$ is a morphism of 2-Hilbert spaces that is also a symmetric monoidal functor. \end{defn} Recall that a functor $F \colon C \to C'$ between monoidal categories is monoidal if it is equipped with a natural isomorphism $\Phi_{x,y} \colon F(x) \otimes F(y) \to F(x \otimes y)$ making the following diagram commute for any objects $x,y,z \in C$: \[ \begin{diagram} [F(x) \otimes (F(y) \otimes F(z))] \node{(F(x) \otimes F(y)) \otimes F(z)} \arrow{e,t}{\Phi_{x,y}\, \otimes 1_{F(z)}} \arrow{s,l}{a_{F(x),F(y),F(z)}} \node{F(x \otimes y) \otimes F(z)} \arrow{e,t}{\Phi_{x\otimes y,z}} \node{F((x \otimes y)\otimes z)} \arrow{s,r}{F(a_{x,y,z})} \\ \node{F(x) \otimes (F(y) \otimes F(z))} \arrow{e,t}{1_{F(x)} \otimes \Phi_{y,z}} \node{F(x) \otimes F(y \otimes z)} \arrow{e,t}{\Phi_{x,y\otimes z}} \node{F(x \otimes (y \otimes z))} \end{diagram} \] together with an isomorphism $\phi \colon 1_{C'} \to F(1_{C})$ making the following diagrams commute for any object $x \in C$: \[ \begin{diagram}[F(1) \otimes F(x)] \node{1 \otimes F(x)} \arrow{e,t}{\ell_{F(x)}} \arrow{s,l}{\phi \otimes 1_{F(x)}} \node{F(x)} \\ \node{F(1) \otimes F(x)} \arrow{e,t}{\Phi_{1,x}} \node{F(1 \otimes x)} \arrow{n,r}{F(\ell_x)} \end{diagram} \] \[ \begin{diagram}[F(1) \otimes F(x)] \node{F(x) \otimes 1} \arrow{e,t}{r_{F(x)}} \arrow{s,l}{1_{F(x)}\otimes \phi} \node{F(x)} \\ \node{F(x) \otimes F(1)} \arrow{e,t}{\Phi_{x,1}} \node{F(x \otimes 1)} \arrow{n,r}{F(r_x)} \end{diagram} \] If $C$ and $C'$ are braided, we say that $F$ is braided if additionally it makes the following diagram commute for all $x,y \in C$: \[ \begin{diagram}[F(x) \otimes F(y)] \node{F(x) \otimes F(y)} \arrow{e,t}{B_{F(x),F(y)}} \arrow{s,l}{\Phi_{x,y}} \node{F(y) \otimes F(x)} \arrow{s,r}{\Phi_{y,x}} \\ \node{F(x \otimes y)} \arrow{e,t}{F(B_{x,y})} \node{F(y \otimes x)} \end{diagram} \] A symmetric monoidal functor is simply a braided monoidal functor that happens to go between symmetric monoidal categories! No extra condition is involved here. Note that if $F \colon H \to H'$ is a homomorphism of braided 2-H*-algebras, $F$ maps any well-balanced adjunction in $H$ to one in $H'$. Thus it preserves dimension in the following sense: \[ \dim(F(x)) = F(\dim(x)) \] for any object $x \in H$. In particular, if $H$ and $H'$ are connected, so that we can identify the dimension of objects in either with numbers, we have simply $\dim(F(x)) = \dim(x)$. \begin{defn}\hspace{-0.08in}{\bf .}\hspace{0.1in} If $H$ and $H'$ are 2-H*-algebras, possibly braided or symmetric, and $F,G \colon H \to H'$ are homomorphisms of the appropriate sort, a {\rm 2-homomorphism} $\alpha \colon F \Rightarrow G$ is a monoidal natural transformation. \end{defn} Suppose that the $(F,\Phi,\phi)$ and $(G,\Gamma,\gamma)$ are monoidal functors from the monoidal category $C$ to the monoidal category $D$. Then a natural transformation $\alpha \colon F \to G$ is monoidal if the diagrams \[ \begin{diagram}[F(x) \otimes F(y)] \node{F(x) \otimes F(y)} \arrow{e,t} {\alpha_x \otimes \alpha_y} \arrow{s,l}{\Phi_{x,y}} \node{G(x) \otimes G(y)} \arrow{s,r}{\Gamma_{x,y}} \\ \node{F(x \otimes y)} \arrow{e,t}{\alpha_{x\otimes y}} \node{G(x \otimes y)} \end{diagram} \] and \[ \begin{diagram}[F(1)] \node{1} \arrow{s,l}{\phi} \arrow{se,t}{\gamma} \\ \node{F(1)} \arrow{e,t}{\alpha_1} \node{G(1)} \end{diagram} \] commute. There are no extra conditions required of `braided monoidal' or `symmetric monoidal' natural transformations. Finally, when we speak of two 2-H*-algebras $H$ and $H'$, possibly braided or symmetric, being {\it equivalent}, we always mean the existence of homomorphisms $F \colon H \to H'$ and $G \colon H' \to H$ of the appropriate sort that are inverses up to a 2-isomorphism. \section{Reconstruction Theorems} \label{recon} In this section we give a classification of symmetric 2-H*-algebras. Doplicher and Roberts proved a theorem which implies that connected even symmetric 2-H*-algebras are all equivalent to categories of compact group representations \cite{DR,DR2}. Here and in all that follows, by a `representation' of a compact group we mean a finite-dimensional continuous unitary representation. Given a compact group $G$, let ${\rm Rep}(G)$ denote the category of such representations of $G$. This becomes a connected even symmetric 2-H*-algebra in an obvious way. While Doplicher and Roberts worked using the language of `C*-categories', their result can be stated as follows: \begin{thm}\label{dhr1}\hspace{-0.08in}{\bf .}\hspace{0.1in} {\rm (Doplicher-Roberts)} Let $H$ be a connected even symmetric 2-H*-\break algebra. Then there exists a homomorphism of symmetric 2-H*-algebras $T \colon H \to {\rm Hilb}$, unique up to a unitary 2-homomorphism. Let $U(T)$ be the group of unitary 2-homomorphisms $\alpha \colon T \Rightarrow T$, given the topology in which a net $\alpha_\lambda \in U(T)$ converges to $\alpha$ if and only if $(\alpha_\lambda)_x \to \alpha_x$ in norm for all $x \in H$. Then $U(T)$ is compact, each Hilbert space $T(x)$ becomes a representation of $U(T)$, and the resulting homomorphism $\tilde T \colon H \to {\rm Rep}(U(T))$ extends to an equivalence of symmetric 2-H*-algebras. \end{thm} Note that any continuous homomorphism $\rho \colon G \to G'$ between compact groups determines a homomorphism of symmetric 2-H*-algebras, \[ \rho^\ast \colon {\rm Rep}(G') \to {\rm Rep}(G), \] sending each representation $\sigma$ of $G'$ to the representation $\sigma \circ \rho$ of $G$. The above theorem yields a useful converse to this construction: \begin{cor}\label{dhr2}\hspace{-0.08in}{\bf .}\hspace{0.1in} {\rm (Doplicher-Roberts)} Let $F \colon H' \to H$ be a homomorphism of connected even symmetric 2-H*-algebras. Let $T \colon H \to {\rm Hilb}$ be a homomorphism of symmetric 2-H*-algebras. Then there exists a continuous group homomorphism \[ F^\ast \colon U(T) \to U(FT) \] such that $F^\ast(\alpha)$ equals the horizontal composite $F\circ \alpha$. Moreover, $(F^\ast)^\ast$ equals $F$ up to a unitary 2-homomorphism. \end{cor} Dolan \cite{Dolan} has noted that a generalization of the Doplicher-Roberts theorem to even symmetric 2-H*-algebras --- not necessarily connected --- amounts to a categorification of the Gelfand-Naimark theorem. The spectrum of a commutative H*-algebra $H$ is a set ${\rm Spec}(H)$ whose points are homomorphisms from $H$ to ${\Bbb C}$. The Gelfand-Naimark theorem implies that $H$ is isomorphic to the algebra of functions from ${\rm Spec}(H)$ to ${\Bbb C}$. Similarly, we may define the `spectrum' of an even symmetric 2-H*-algebra $H$ to be the groupoid ${\rm Spec}(H)$ whose objects are homomorphisms from $H$ to ${\rm Hilb}$, and whose morphisms are unitary 2-homomorphisms between these. Moreover, we shall show that $H$ is equivalent to a symmetric 2-H*-algebra whose objects are `representations' of ${\rm Spec}(H)$ --- certain functors from ${\rm Spec}(H)$ to ${\rm Hilb}$. Indeed, our proof of this uses an equivalence \[ \hat{\hbox{\hskip 0.5em}}\colon H \to {\rm Rep}({\rm Spec}(H)) \] that is just the categorified version of the `Gelfand transform' for commutative H*-algebras. In fact, there is no need to restrict ourselves to symmetric 2-H*-algebras that are even. To treat a general symmetric 2-H*-algebra $H$ we need objects of ${\rm Spec}(H)$ to be homomorphisms from $H$ to a symmetric 2-H*-algebra of `super-Hilbert spaces'. The spectrum will then be a `supergroupoid' --- though not the most general sort of thing one could imagine calling a supergroupoid. \begin{defn} \hspace{-0.08in}{\bf .}\hspace{0.1in} Define ${\rm SuperHilb}$ to be the category whose objects are $Z_2$-graded (finite-dimensional) Hilbert spaces, and whose morphisms are linear maps preserving the grading. \end{defn} The category ${\rm SuperHilb}$ can be made into a symmetric 2-H*-algebra where the $\ast$-structure is the ordinary Hilbert space adjoint, the product is the usual tensor product of ${\Bbb Z}_2$-graded Hilbert spaces, and the braiding is given on homogeneous elements $v \in x$, $w \in y$ by \[ B_{x,y}(v \otimes w) = (-1)^{{\rm deg} v\, {\rm \deg} w} \, w \otimes v. \] \begin{defn} \hspace{-0.08in}{\bf .}\hspace{0.1in} If $H$ is a symmetric 2-H*-algebra, define ${\rm Spec}(H)$ to be the category whose objects are symmetric 2-H*-algebra homomorphisms $F \colon H \to {\rm SuperHilb}$ and whose morphisms are unitary 2-homomorphisms between these. \end{defn} \begin{defn} \hspace{-0.08in}{\bf .}\hspace{0.1in} A {\rm topological groupoid} is a groupoid for which the $\hom$-sets are topological spaces and the groupoid operations are continuous. A {\rm compact groupoid} is is a topological groupoid with compact Hausdorff $\hom$-sets and finitely many isomorphism classes of objects. \end{defn} \begin{defn} \hspace{-0.08in}{\bf .}\hspace{0.1in} A {\rm supergroupoid} is a groupoid $G$ equipped with a natural transformation $\beta \colon 1_G \Rightarrow 1_G$, the {\rm balancing,} with $\beta^2 = 1$. A {\rm compact supergroupoid} is a supergroupoid that is also a compact groupoid. \end{defn} Let $H$ be a symmetric 2-H*-algebra. Then ${\rm Spec}(H)$ becomes a topological groupoid if for any $S,T \colon H \to {\rm Hilb}$ we give $\hom(S,T)$ the topology in which a net $\alpha_\lambda$ converges to $\alpha$ if and only if $(\alpha_\lambda)_x \to \alpha_x$ in norm for any $x \in H$. We shall show that ${\rm Spec}(H)$ is a compact groupoid. Also, ${\rm Spec}(H)$ becomes a supergroupoid if for any object $T \in {\rm Spec}(H)$ we define $\beta_T \colon T \Rightarrow T$ by \[ (\beta_T)_x = b_{T(x)} = T(b_x) \] for any object $x \in H$. One can check that $\beta \colon 1_{{\rm Spec}(H)} \Rightarrow 1_{{\rm Spec}(H)}$ is a natural transformation, and $\beta^2 = 1$ because the balancing for $H$ satisfies $b_x^2 = 1$ for any $x \in H$. \begin{defn}\hspace{-0.08in}{\bf .}\hspace{0.1in} Given a compact supergroupoid $G$, a (continuous, unitary, finite-dimensional) {\rm representation} of $G$ is a functor $F \colon G \to {\rm SuperHilb}$ such that $F(g)$ is unitary for every morphism $g$ in $G$, $F \colon \hom(x,y) \to \hom(F(x),F(y))$ is continuous for all objects $x,y \in G$, and \[ F(\beta_x) = b_{F(x)} \] for every object $x \in G$. We define ${\rm Rep}(G)$ to be the category having representations of $G$ as objects and natural transformations between these as morphisms. \end{defn} Let $G$ be a compact supergroupoid. Then the category ${\rm Rep}(G)$ becomes an even symmetric 2-H*-algebra in a more or less obvious way as follows. Given objects $F,F' \in {\rm Rep}(G)$, we make $\hom(F,F')$ into a Hilbert space with the obvious linear structure and the inner product given by \[ \langle \alpha, \beta \rangle = \sum_x {\rm tr}({\alpha_x}^\ast \beta_x) \] where the sum is taken over any maximal set of nonisomorphic objects of $G$. This makes ${\rm Rep}(G)$ into a ${\rm Hilb}$-category. Moreover, ${\rm Rep}(G)$ becomes a 2-Hilbert space if we define the dual of $\alpha \colon F \Rightarrow F'$ by $(\alpha^\ast)_x = (\alpha_x)^\ast$. We define the tensor product of objects $F,F' \in {\rm Rep}(G)$ by \[ (F \otimes F')(x) = F(x) \otimes F'(x), \qquad (F \otimes F')(f) = F(f) \otimes F'(f) \] for any object $x \in G$ and morphism $f$ in $G$. It is easy to define a tensor product of morphisms and associator making ${\rm Rep}(G)$ into a monoidal category, and to check that ${\rm Rep}(G)$ is then a 2-H*-algebra. Finally, ${\rm Rep}(G)$ inherits a braiding from the braiding in ${\rm SuperHilb}$, making ${\rm Rep}(G)$ into a symmetric 2-H*-algebra. Now suppose $H$ is an even symmetric 2-H*-algebra. Then there is a functor \[ \hat{\hbox{\hskip 0.5em}}\colon H \to {\rm Rep}({\rm Spec}(H)) , \] the {\it categorified Gelfand transform}, given as follows. For every object $x \in H$, $\hat x$ is the representation with \[ \hat x(T) = T(x) \] for all $T \in {\rm Spec}(H)$, and \[ \hat x(\alpha) = \alpha_x \] for all $\alpha \colon T \Rightarrow T'$, where $T,T' \in {\rm Spec}(H)$. For every morphism $f \colon x \to y$ in $H$, $\hat f \colon \hat x \Rightarrow \hat y$ is the natural transformation with \[ \hat f(T) = T(f) \] for all $T \in {\rm Spec}(H)$. Our generalized Doplicher-Roberts theorem states: \begin{thm} \label{dhr4}\hspace{-0.08in}{\bf .}\hspace{0.1in} Suppose that $H$ is a symmetric 2-H*-algebra. Then ${\rm Spec}(H)$ is a compact supergroupoid and $\hat{\hbox{\hskip 0.5em}}\colon H \to {\rm Rep}({\rm Spec}(H))$ extends to an equivalence of symmetric 2-H*-algebras. \end{thm} Proof - We have described how ${\rm Spec}(H)$ is a supergroupoid. To see that it is compact, note that for any $S,T \in {\rm Spec}(H)$ the $\hom$-set $\hom(S,T)$ is a compact Hausdorff space, by Tychonoff's theorem. We also need to show that ${\rm Spec}(H)$ has finitely many isomorphism classes of objects. The unit object $1_H$ is the direct sum of finitely many nonisomorphic simple objects $e_i$, the kernels of the minimal projections $p_i$ in the commutative H*-algebra ${\rm end}(1_H)$. Any object $x \in H$ is thus a direct sum of objects $x_i = e_i \otimes x$, and any morphism $f \colon x \to y$ is a direct sum of morphisms $f_i \colon x_i \to y_i$. In short, $H$ is, in a fairly obvious sense, the direct sum of finitely many connected symmetric 2-H*-algebras $H_i$. Any homomorphism $T \colon H \to {\rm SuperHilb}$ induces a homomorphism from ${\rm end}(1_H)$ to ${\rm end}(1_{{\rm SuperHilb}}) \cong {\Bbb C}$, which must annihilate all but one of the projections $p_i$, so $T$ sends one of the objects $x_i$ to $1_{{\rm SuperHilb}}$ and the rest to $0$. Thus ${\rm Spec}(H)$ is, as a groupoid, equivalent to the disjoint union of the groupoids ${\rm Spec}(H_i)$, and hence has finitely many isomorphism classes of objects. To show that the categorified Gelfand transform is an equivalence, first suppose that $H$ is even and connected. Then the supergroupoid ${\rm Spec}(H)$ has $\beta = 1$, so every representation $F \colon {\rm Spec}(H) \to {\rm SuperHilb}$ factors through the inclusion ${\rm Hilb} \hookrightarrow {\rm SuperHilb}$. Moreover, by Theorem \ref{dhr1} all the objects of ${\rm Spec}(H)$ are isomorphic, so ${\rm Spec}(H)$ is equivalent, as a groupoid, to the group ${\rm U}(T)$ for any $T \in {\rm Spec}(H)$. We thus obtain an equivalence of symmetric 2-H*-algebras between ${\rm Rep}({\rm Spec}(H))$ and ${\rm Rep}({\rm U}(T))$ as defined in Theorem \ref{dhr1}. Using this, the fact that $\tilde T \colon H \to {\rm Rep}({\rm U}(T))$ is an equivalence translates into the fact that $\hat{\hbox{\hskip 0.5em}}\colon H \to {\rm Rep}({\rm Spec}(H))$ is an equivalence. Next, suppose that $H$ is even but not connected. Then $H$ is a direct sum of the even connected symmetric 2-H*-algebras $H_i$ as above, and ${\rm Rep}({\rm Spec}(H))$ is similarly the direct sum of the ${\rm Rep}({\rm Spec}(H_i))$. Because the categorified Gelfand transform $\hat{\hbox{\hskip 0.5em}}\colon H_i \to {\rm Rep}({\rm Spec}(H_i))$ is an equivalence for all $i$, $\hat{\hbox{\hskip 0.5em}}\colon H \to {\rm Rep}({\rm Spec}(H))$ is an equivalence. Finally we treat the general case where $H$ is an arbitrary symmetric 2-H*-algebra. Note that if $H$ and $K$ are symmetric 2-H*-algebras, a symmetric 2-H*-algebra homomorphism $F \colon H \to K$ gives rise to a symmetric 2-H*-algebra homomorphism $F^\flat \colon H^\flat \to K^\flat$ between their bosonizations, where $F^\flat$ is the same as $F$ on objects and morphisms. Note also that $F$ is an equivalence of symmetric 2-H*-algebras if and only if $F^\flat$ is. Thus to show that $\hat{\hbox{\hskip 0.5em}}\colon H \to {\rm Rep}({\rm Spec}(H))$ is an equivalence, it suffices to show $\hat{\hbox{\hskip 0.5em}}^\flat \colon H^\flat \to {\rm Rep}({\rm Spec}(H))^\flat$ is an equivalence. For this, note that any supergroupoid $G$ has a {\it bosonization} $G^\flat$, in which the underlying compact groupoid of $G$ is equipped with the trivial balancing $\beta = 1$. Moreover, there is a homomorphism of symmetric 2-H*-algebras \[ \begin{diagram}[{\rm Rep}(G)^\flat] \node{{\rm Rep}(G)^\flat} \arrow{e,t}{X} \node{{\rm Rep}(G^\flat)} \end{diagram} \] sending any representation $F \in {\rm Rep}(G)^\flat$ to the representation $X(F) \in {\rm Rep}(G^\flat)$ given by the commutative square \[ \begin{diagram}[{\rm SuperHilb}] \node{G^\flat} \arrow{e,t}{X(F)} \arrow{s,l}{I} \node{{\rm SuperHilb}} \\ \node{G} \arrow{e,t}{F} \node{{\rm SuperHilb}} \arrow{n,r}{E} \end{diagram} \] Here $I \colon G^\flat \to G$ is the identity on the underlying groupoids, while the 2-H*-algebra homomorphism $E \colon {\rm SuperHilb} \to {\rm SuperHilb}$ maps any super-Hilbert space to the even super-Hilbert space with the same underlying Hilbert space, and acts as the identity on morphisms. One may check that $X(F)$ is really a compact supergroupoid representation. Similarly, given a morphism $\alpha \colon F \Rightarrow F'$ in ${\rm Rep}(G)^\flat$, we define $X(\alpha)$ to be the horizontal composite $I \circ \alpha \circ E$. In fact, $X$ is an equivalence, for given any representation $F$ of $G^\flat$ we can turn it back into a representation of $G$ by equipping each Hilbert space $F(x)$, $x \in G$ with the grading $F(\beta_x)$, where $\beta$ is the balancing of $G$. Similarly, for any symmetric 2-H*-algebra $H$ there is an equivalence \[ \begin{diagram}[{\rm Spec}(H)^\flat] \node{{\rm Spec}(H)^\flat} \arrow{e,t}{Y} \node{{\rm Spec}(H^\flat)} \end{diagram} \] sending any object $T \in {\rm Spec}(H)^\flat$ to the object $Y(T) \in {\rm Spec}(H^\flat)$ given by the commutative square \[ \begin{diagram}[{\rm SuperHilb}] \node{H^\flat} \arrow{e,t}{Y(T)} \arrow{s,l}{I} \node{{\rm SuperHilb}} \\ \node{H} \arrow{e,t}{T} \node{{\rm SuperHilb}} \arrow{n,r}{E} \end{diagram} \] where $I \colon H^\flat \to H$ is the identity on the underlying 2-H*-algebras, while $E$ is given as above. We thus have equivalences \[ \begin{diagram}[{\rm Rep}({\rm Spec}(H))^\flat] \node{{\rm Rep}({\rm Spec}(H))^\flat} \arrow{e,t}{\sim} \node{{\rm Rep}({\rm Spec}(H)^\flat)} \arrow{e,t}{\sim} \node{{\rm Rep}({\rm Spec}(H^\flat))} \end{diagram} \] and their composite gives a diagram commuting up to natural isomorphism: \[ \begin{diagram}[{\rm Rep}({\rm Spec}(H))^\flat] \node{H^\flat} \arrow{e,t}{\hat{\hbox{\hskip 0.5em}}^\flat} \arrow{se,b}{\hat{\hbox{\hskip 0.5em}}} \node{{\rm Rep}({\rm Spec}(H))^\flat} \arrow{s}\\ \node[2]{{\rm Rep}({\rm Spec}(H^\flat))} \end{diagram} \] It follows that $\hat{\hbox{\hskip 0.5em}}^\flat \colon H^\flat \to {\rm Rep}({\rm Spec}(H))^\flat$ is an equivalence, as was to be shown. \hbox{\hskip 30em} \hskip 3em \hbox{\BOX} \vskip 2ex Presumably what Theorem \ref{dhr4} is trying to tell us is that there are 2-functors ${\rm Rep}$ and ${\rm Spec}$ going both ways between the 2-category of compact supergroupoids and the 2-category of symmetric 2-H*-algebras, and that these extend to a 2-equivalence of 2-categories. We shall not try to prove this here. However, it is worth noting that for any compact supergroupoid $G$, there is a functor \[ \check{\hbox{\hskip 0.5em}} \colon G \to {\rm Spec}({\rm Rep}(G)) \] given as follows. For every object $x \in G$, $\hat x$ is the object of ${\rm Spec}({\rm Rep}(G))$ with \[ \hat x(F) = F(x) \] for all $F \in {\rm Rep}(G)$, and \[ \hat x(\alpha) = \alpha_x \] for all $\alpha \colon F \to F'$, where $F,F' \in {\rm Rep}(G)$. For every morphism $g \colon x \to y$ in $G$, $\check g \colon \check x \Rightarrow \check y$ is the natural transformation with \[ \check g(F) = F(g) \] for all $F \in {\rm Rep}(G)$. Presumably $\check{\hbox{\hskip 0.5em}} \colon G \to {\rm Spec}({\rm Rep}(G))$ is in some sense an equivalence of compact supergroupoids. \subsection{Compact abelian groups} The representation theory of compact abelian groups is rendered especially simple by the use of Fourier analysis, as generalized by Pontryagin. Suppose that $T$ is a compact abelian group. Then its dual $\hat T$ is defined as the set of equivalence classes of irreducible representations $\rho$ of $T$. The dual becomes a discrete abelian group with operations given as follows: \[ [\rho][\rho'] = [\rho \otimes \rho'], \] \[ [\rho]^{-1} = [\rho^\ast] \] Then the Fourier transform is a unitary isomorphism \[ f \colon L^2(T) \to L^2(\hat T) \] given by \[ f(\chi_\rho) = \delta_{[\rho]} \] where $\chi_\rho$ is the character of the representation $\rho$, and $\delta_{[\rho]}$ is the function on $\hat T$ which equals $1$ at $[\rho]$ and $0$ elsewhere. The Fourier transform has an interesting categorification. Note that the ordinary Fourier transform has as its domain the infinite-dimensional Hilbert space $L^2(T)$, which has a basis given by the characters of irreducible representations of $T$. The categorified Fourier transform will have as domain the 2-Hilbert space ${\rm Rep}(T)$, which has a basis given by the irreducible representations themselves. (Taking the character of a representation is a form of `decategorification'.) Similarly, just as the ordinary Fourier transform has as its codomain an infinite-dimensional Hilbert space of ${\Bbb C}$-valued functions on $\hat T$, the categorified Fourier transform will have as its codomain a 2-Hilbert space of ${\rm Hilb}$-valued functions on $\hat T$. More precisely, define ${\rm Hilb}[G]$ for any discrete group $G$ to be the category whose objects are $G$-graded Hilbert spaces for which the total dimension is finite, and whose morphisms are linear maps preserving the grading. Alternatively, we can think of ${\rm Hilb}[G]$ as the category of hermitian vector bundles over $G$ for which the sum of the dimensions of the fibers is finite. We may write any object $x \in {\rm Hilb}[G]$ as a $G$-tuple $\{x(g)\}_{g \in G}$ of Hilbert spaces. The category ${\rm Hilb}[G]$ becomes a 2-H*-algebra in an obvious way with a product modelled after the convolution product in the group algebra ${\Bbb C}[G]$: \[ (x \otimes y)(g) = \bigoplus_{\{g',g'' \in G\,\colon\; g'g'' = g\}} x(g') \otimes y(g'') .\] If $G$ is abelian, ${\rm Hilb}[G]$ becomes a symmetric 2-H*-algebra. Now suppose that $T$ is a compact abelian group. Given any object $x \in {\rm Rep}(T)$, we may decompose $x$ into subspaces corresponding to the irreducible representations of $T$: \[ x = \bigoplus_{g \in \hat T} x(g) . \] We define the {\it categorified Fourier transform} \[ F \colon {\rm Rep}(T) \to {\rm Hilb}[\hat T] \] as follows. For any object $x \in {\rm Rep}(T)$, we set \[ F(x) = \{x(g)\}_{g \in \hat T} . \] Moreoever, any morphism $f \colon x \to y$ in ${\rm Rep}(T)$ gives rise to linear maps $f(g) \colon x(g) \to y(g)$ and thus a morphism $F(f)$ in ${\rm Hilb}[\hat T]$. One can check that $F$ is not only 2-Hilbert space morphism but actually a homomorphism of symmetric 2-H*-algebras. This is the categorified analog of how the ordinary Fourier transform sends pointwise multiplication to convolution. Note that, in analogy to the formula \[ f(\chi_\rho) = \delta_{[\rho]} \] satisfied by the ordinary Fourier transform, for any irreducible representation $\rho$ of $T$ the categorified Fourier transform $F(\rho)$ is a hermitian vector bundle that is 1-dimensional at $[\rho]$ and 0-dimensional elsewhere. \begin{thm} \hspace{-0.08in}{\bf .}\hspace{0.1in} If $T$ is a compact abelian group, the categorified Fourier transform $F \colon {\rm Rep}(T) \to {\rm Hilb}(\hat T)$ is an equivalence of symmetric 2-H*-algebras. \end{thm} Proof - There is a homomorphism $G \colon {\rm Hilb}[\hat T] \to {\rm Rep}(T)$ sending each object $\{x(g)\}_{g \in \hat T}$ in ${\rm Hilb}[\hat T]$ to a representation of $T$ which is a direct sum of spaces $x(g)$ transforming according to the different isomorphism classes $g \in \hat T$ of irreducible representations of $T$. One can check that $FG$ and $GF$ are naturally isomorphic to the identity. \hskip 3em \hbox{\BOX} \vskip 2ex \subsection{Compact classical groups} \label{classical} The representation theory of a `classical' compact Lie group has a different flavor from that of general compact Lie groups. The representation theory of general compact Lie groups heavily involves the notions of maximal torus, Weyl group, roots and weights. We hope to interpret this theory in terms of 2-Hilbert spaces in a future paper. However, the representation theory of a classical group can also be studied using Young diagrams \cite{Weyl}. This approach relies on the fact that its categories of representations have simple universal properties. These universal properties can be described in the language of symmetric 2-H*-algebras, and a description along these lines represents a distilled version of the Young diagram theory. For example, consider the group ${\rm U}(n)$. The fundamental representation of ${\rm U}(n)$ on ${\Bbb C}^n$ is the `universal $n$-dimensional representation'. In other words, for a group to have a (unitary) representation on ${\Bbb C}^n$ is precisely for it to have a homomorphism to ${\rm U}(n)$. This universal property can also be expressed as a universal property of ${\rm Rep}({\rm U}(n))$. Suppose that $G$ is a compact group. Then any $n$-dimensional representation $y \in {\rm Rep}(G)$ is isomorphic to a representation of the form $\rho \colon G \to {\rm U}(n)$. The representation $\rho$ gives rise to a homomorphism \[ \rho^\ast \colon {\rm Rep}({\rm U}(n)) \to {\rm Rep}(G), \] and letting $x$ denote the fundamental representation of ${\rm U}(n)$, we have $\rho^\ast(x) = \rho$. Since $\rho$ and $y$ are isomorphic, there is a unitary 2-homomorphism from $\rho^\ast$ to a homomorphism \[ F \colon {\rm Rep}({\rm U}(n)) \to {\rm Rep}(G) \] with $F(x) = y$. In short, for any $n$-dimensional object $y \in {\rm Rep}(G)$ there is a homomorphism $F \colon {\rm Rep}({\rm U}(n)) \to {\rm Rep}(G)$ of symmetric 2-H*-algebras with $F(x) = y$. On the other hand, suppose $F' \colon {\rm Rep}({\rm U}(n)) \to {\rm Rep}(G)$ is any other homomorphism with $F'(x) = y$. We claim that there is a unitary 2-homomorphism from $F$ to $F'$. By Corollary \ref{dhr2}, there exists a homomorphism $\rho' \colon G \to {\rm U}(n)$ with a unitary 2-homomorphism from $F'$ to $\rho'^\ast$. On the other hand, by construction there is a unitary 2-homomorphism from $F$ to $\rho^\ast$ for some $\rho \colon G \to {\rm U}(n)$. To show there is a unitary 2-homomorphism from $F$ to $F'$, it thus suffices to show that $\rho$ and $\rho'$ are isomorphic in ${\rm Rep}(G)$. This holds because $\rho \cong y = F'(x) \cong \rho'^\ast(x) = \rho'$. Now, since any connected even symmetric 2-H*-algebra is unitarily equivalent to ${\rm Rep}(G)$ for some compact $G$ by Theorem \ref{dhr1}, we may restate these results as follows. Suppose $H$ is a connected even symmetric 2-H*-algebra and let $y$ be an $n$-dimensional object of $H$. Then there exists a homomorphism $F \colon {\rm Rep}({\rm U}(n)) \to H$ with $F(x) = y$. Moreover, this is unique up to a unitary 2-homomorphism. Furthermore, we can drop the assumption that $H$ is even by working with the full subcategory whose objects are all the even objects of $H$. We may thus state the universal property of ${\rm Rep}({\rm U}(n))$ as follows: \begin{thm}\label{un}\hspace{-0.08in}{\bf .}\hspace{0.1in} ${\rm Rep}({\rm U}(n))$ is the free connected symmetric 2-H*-algebra on an even object $x$ of dimension $n$. That is, given any even $n$-dimensional object $y$ of a connected symmetric 2-H*-algebra $H$, there exists a homomorphism of symmetric 2-H*-algebras $F \colon {\rm Rep}(U(n)) \to H$ with $F(x) = y$, and $F$ is unique up to a unitary 2-homomorphism. \end{thm} Let $\Lambda^n x$ denote the cokernel of $p_A \colon x^{\otimes n} \to x^{\otimes n}$ (complete antisymmetrization), and let $S^n x$ denote the cokernel of $p_S \colon x^{\otimes n} \to x^{\otimes n}$ (complete symmetrization). We can describe the category of representations of ${\rm SU}(n)$ as follows: \begin{thm}\label{sun} \hspace{-0.08in}{\bf .}\hspace{0.1in} ${\rm Rep}({\rm SU}(n))$ is the free connected symmetric 2-H*-algebra on an even object $x$ with $\Lambda^n x \cong 1$. \end{thm} Proof - Suppose that $G$ is a compact group and the object $y \in {\rm Rep}(G)$ has $\Lambda^n y \cong 1$. It follows that $y$ is $n$-dimensional by the computation in Proposition \ref{dim}, and the isomorphism $\Lambda^n y \cong 1$ determines a $G$-invariant volume form on the representation $y$. Thus $y$ is isomorphic to a representation of the form $\rho \colon G \to {\rm SU}(n)$. The rest of the proof follows that of Theorem \ref{un}. \hskip 3em \hbox{\BOX} \vskip 2ex Here we can see in a simple context how our theory is a distillation of the theory of Young diagrams. (The Young diagram approach to representation theory is more familiar for ${\rm SU}(n)$ than for ${\rm U}(n)$.) In heuristic terms, the above theorem says that every representation of ${\rm SU}(n)$ is generated from the fundamental representation $x$ using the operations present in a symmetric 2-H*-algebra --- the $\ast$-structure, direct sums, cokernels, tensor products, duals, and the symmetry --- with no relations other than those implied by the axioms for a connected symmetric 2-H*-algebra and the fact that $x$ is even and $\Lambda^n x \cong 1$. The theory of Young diagrams makes this explicit by listing the irreducible representations of ${\rm SU}(n)$ in terms of minimal projections $p \colon x^{\otimes x} \to x^{\otimes k}$, or in other words, Young diagrams with $k$ boxes. The symmetric 2-H*-algebra of representations of a subgroup $G \subset {\rm SU}(n)$, such as ${\rm SO}(n)$ or ${\rm Sp}(n)$, is a quotient of ${\rm Rep}({\rm SU}(n))$. We may describe this quotienting process by giving extra relations as in Theorems \ref{spn} and \ref{son} below. These extra relations give identities saying that different Young diagrams correspond to the same representation of $G$. The classical groups $\O(n)$ and ${\rm Sp}(n)$ are related to the concept of self-duality. Given adjunctions $(x,x^\ast,i_x,e_x)$ and $(y,y^\ast,i_y,e_y)$ in a monoidal category $C$, for any morphism $f \colon x \to y$ there is a morphism $f^\dagger \colon y^\ast \to x^\ast$, given in the strict case by the composite: \[ \begin{diagram}[y^\ast \otimes x \otimes x^\ast] \node{y^\ast = y^\ast \otimes 1} \arrow{e,t}{y^\ast \otimes i_x} \node{y^\ast \otimes x \otimes x^\ast} \arrow{e,t}{y^\ast \otimes f \otimes x^\ast} \node{y^\ast \otimes y \otimes x^\ast} \arrow{e,t}{e_y \otimes x^\ast} \node{1 \otimes x^\ast = x^\ast} \end{diagram} \] (Our notation here differs from that of HDA0.) Since the left dual of an object in a 2-Hilbert space is also its right dual as in Proposition \ref{2H*2}, given a morphism $f \colon x \to x^\ast$ we obtain another morphism $f^\dagger \colon x \to x^\ast$. Using this we may describe ${\rm Rep}(\O(n))$ and ${\rm Rep}({\rm Sp}(n))$ as certain `free connected symmetric 2-H*-algebras on one self-dual object': \begin{thm}\label{on} \hspace{-0.08in}{\bf .}\hspace{0.1in} ${\rm Rep}(\O(n))$ is the free connected symmetric 2-H*-algebra on an even object $x$ of dimension $n$ with an isomorphism $f \colon x \to x^\ast$ such that $f^\dagger = f$. \end{thm} Proof - Suppose that $G$ is a compact group and the object $y \in {\rm Rep}(G)$ is $n$-dimensional and equipped with an isomorphism $f \colon x \to x^\ast$ with $f^\dagger = f$. Then there is a nondegenerate pairing $F \colon y \otimes y \to 1$ given by $F = (y \otimes f)i_y^\ast$. A calculation, given in the proof of Proposition \ref{selfdual}, shows that $F$ is symmetric. It follows that $y$ is isomorphic to a representation of the form $\rho \colon G \to \O(n)$. The rest of the proof follows that of Theorem \ref{un}. \hskip 3em \hbox{\BOX} \vskip 2ex \begin{thm}\label{spn} \hspace{-0.08in}{\bf .}\hspace{0.1in} ${\rm Rep}({\rm Sp}(n))$ is the free connected symmetric 2-H*-algebra on one even object $x$ with $\Lambda^n x \cong 1$ and with an isomorphism $f \colon x \to x^\ast$ such that $f^\dagger = -f$. \end{thm} Proof - The proof is analogous to that of Theorem \ref{on}, except that the pairing $F$ is skew-symmetric. \hskip 3em \hbox{\BOX} \vskip 2ex Following the proof of Theorem \ref{sun} we may also characterize ${\rm Rep}({\rm SO}(n))$ as follows: \begin{thm}\label{son} \hspace{-0.08in}{\bf .}\hspace{0.1in} ${\rm Rep}({\rm SO}(n))$ is the free connected symmetric 2-H*-algebra on an even object $x$ with $\Lambda^n x \cong 1$ and with an isomorphism $f \colon x \to x^\ast$ such that $f^\dagger = f$. \end{thm} The conditions on the isomorphism $f \colon x \to x^\ast$ in Theorems \ref{on} and \ref{spn} are quite reasonable, in the following sense: \begin{prop}\label{selfdual} \hspace{-0.08in}{\bf .}\hspace{0.1in} Suppose that $x$ is a simple object in a symmetric 2-H*-algebra and that $x$ is isomorphic to $x^\ast$. Then one and only one of the following is true: either there is an isomorphism $f \colon x \to x^\ast$ with $f = f^\dagger$, or there is an isomorphism $f \colon x \to x^\ast$ with $f = -f^\dagger$. \end{prop} Proof - Note that there is an isomorphism of complex vector spaces \begin{eqnarray*} \hom(x,x^\ast) &\cong& \hom(x \otimes x,1) \\ f &\mapsto& (1 \otimes f)i_x^\ast, \end{eqnarray*} and note that \[ \hom(x \otimes x,1) \cong \hom(S^2 x,1) \oplus \hom(\Lambda^2 x,1) . \] Suppose $f \colon x \to x^\ast$ is an isomorphism and let $F = (1 \otimes f)i_x^\ast$. Since $x$ is simple, $f$ and thus $F$ is unique up to a scalar multiple, so $F$ must lie either in $\hom(S^2 x,1)$ or $\hom(\Lambda^2 x,1)$. In other words, $B_{x,x}F = \pm F$. Choose a well-balanced adjunction for $x$. Assuming without loss of generality that $H$ is strict, we have \begin{eqnarray*} f^\dagger &=& (x \otimes i_x)(x \otimes f\otimes x^\ast)(i_x^\ast \otimes x^\ast) \\ &=& (x \otimes i_x)(F \otimes x^\ast) \\ &=&\pm (x \otimes i_x)(B_{x,x}F \otimes x^\ast) \\ &=&\pm (f \otimes i_x)(B_{x^\ast,x} \otimes x^\ast)(i_x^\ast \otimes x^\ast)\\ &=& \pm f b_{x^\ast} \end{eqnarray*} Since $b_{x^\ast} = \pm 1_{x^\ast}$ depending on whether $x$, and thus $x^\ast$, is even or odd, we have $f^\dagger = \pm f$. \hskip 3em \hbox{\BOX} \vskip 2ex \noindent This result is well-known if $H$ is a category of compact group representations \cite{FH}. Here one may also think of the morphism $f \colon x \to x^\ast$ as a conjugate-linear intertwining operator $j \colon x \to x$. The condition that $f = \pm f^\dagger$ is then equivalent to the condition that $j^2 = \pm 1_x$. One says that $x$ is a real representation if $j^2 = 1_x$ and a quaternionic representation if $j^2 = -1_x$, establishing the useful correspondence: \[ {\rm real:complex:quaternionic::orthogonal:unitary:symplectic} \] The following alternate characterization of ${\rm Rep}({\rm U}(1))$ is interesting because it emphasizes the relation between duals and inverses. Whenever $T$ is a compact abelian group and $x \in {\rm Rep}(T)$, the dual $x^\ast$ is also the inverse of $x$, in the sense that $x \otimes x^\ast \cong 1$. We have: \begin{thm}\hspace{-0.08in}{\bf .}\hspace{0.1in} ${\rm Rep}({\rm U}(1))$ is the free connected symmetric 2-H*-algebra on an even object $x$ with $x \otimes x^\ast \cong 1$. \end{thm} Proof - By Theorem \ref{un} it suffices to show that an object $x$ in a connected symmetric 2-H*-algebra is 1-dimensional if and only if $x \otimes x^\ast \cong 1$. On the one hand, by the multiplicativity of dimension, $x \otimes x^\ast \cong 1$ implies that $\dim(x) = 1$. On the other hand, suppose $\dim(x) = 1$. Then we claim $i_x \colon 1 \to x \otimes x^\ast$ and $i_x^\ast \colon x \otimes x^\ast \to 1$ are inverses. First, ${i_x}^\ast i_x$ is the identity since $\dim(x) = 1$. Second, ${i_x}^\ast i_x \in {\rm end}(x \otimes x^\ast)$ is idempotent since $\dim(x) = 1$. Since $x \otimes x^\ast$ is 1-dimensional, it is simple (by the additivity of dimension), so ${i_x}^\ast i_x$ must be the identity. \hskip 3em \hbox{\BOX} \vskip 2ex Finally, it is interesting to note that ${\rm SuperHilb}$ is the free connected symmetric 2-H*-algebra on an odd object $x$ with $x \otimes x \cong 1$. This object $x$ is the one-dimensional odd super-Hilbert space. \section{Conclusions} The reader will have noted that some of our results are slight reworkings of those in the literature. One advantage of our approach is that it immediately suggests generalizations to arbitrary $n$. While the general study of $n$-Hilbert spaces will require a deeper understanding of $n$-category theory, we expect many of the same themes to be of interest. With this in mind, let us point out some problems with what we have done so far. One problem concerns the definition of the quantum-theoretic hierarchy. A monoid is a essentially a category with one object. More precisely, a category with one object $x$ can be reconstructed from the monoid ${\rm end}(x)$, and up to isomorphism every monoid comes from a one-object category in this way. Comparing Figures 1 and 2, one might at first hope that by analogy an H*-algebra would be a one-dimensional 2-Hilbert space. Unfortunately, the way we have set things up, this is not the case. If $H$ is a one-dimensional 2-Hilbert space with basis given by the object $x$, then ${\rm end}(x)$ is an H*-algebra. However, ${\rm end}(x)$ is always isomorphic to ${\Bbb C}$; one does not get any other H*-algebras this way. The reason appears to be the requirement that a 2-Hilbert space has cokernels, so that if ${\rm end}(x)$ has nontrivial idempotents, $x$ has subobjects. If we dropped this clause in the definition of a 2-Hilbert space, there would be a correspondence between H*-algebras and 2-Hilbert spaces, all of whose objects are direct sums of a single object $x$. Perhaps in the long run it will be worthwhile to modify the definition of 2-Hilbert space in this way. On the other hand, an H*-algebra is also an H*-category with one object. An H*-category has sums and differences of morphisms, but not of objects, i.e., it need not have direct sums and cokernels. Perhaps, therefore, a $k$-tuply monoidal $n$-Hilbert space should really be some sort of `$(n+k)$-H*-category' with one $j$-morphism for $j < k$, and sums and differences of $j$-morphisms for $j \ge k$. A second problem concerns the program of getting an invariant of $n$-tangles in $(n+k)$-dimensions from an object in a $k$-tuply monoidal $n$-Hilbert space. Let us recall what is known so far here. Oriented tangles in 2 dimensions are the morphisms in a monoidal category with duals, $C_{1,1}$. Here by `monoidal category with duals' we mean a monoidal $\ast$-category in which every object has a left dual, the tensor product is a $\ast$-functor, and the associator is a unitary natural transformation. Suppose that $X$ is any other monoidal category with duals, e.g.\ a 2-H*-algebra. Then any adjunction $(x,x^\ast,i,e)$ in $C$ uniquely determines a monoidal $\ast$-functor $F \colon C_{1,1} \to H$ up to monoidal unitary natural isomorphism. The functor $F$ is determined by the requirement that it maps the positively oriented point to $x$, the negatively oriented point to $x^\ast$, and the appropriately oriented `cup' and `cap' tangles to $e$ and $i$. According to our philosophy we would prefer $F$ to be determined by an object $x \in X$ rather than an adjunction. However $F$ is not determined up to natural transformation by requiring that it map the positively oriented point to $x$. For example, take $X = {\rm Hilb}$ and let $x \in X$ be any object. We may let $F$ send the negatively oriented point to the dual Hilbert space $x^\ast$, and send the cap and cup to the standard linear maps $e \colon x^\ast \otimes x \to {\Bbb C}$ and $i \colon {\Bbb C} \to x \otimes x^\ast$. Then $F(ii^\ast) = \dim(x)1_x$. Alternatively we may let $F$ send the cap and cup to $e' = \alpha^{-1} e$ and $i' = \alpha i$ for any nonzero $\alpha \in {\Bbb C}$. Then $F(ii^\ast) = |\alpha|^2 \dim(x)1_x$. The problem is that while adjunctions in $X$ are unique up to unique isomorphism, the isomorphism is not necessarily unitary. In HDA0 we outlined a way to deal with this problem by `strictifying' the notion of a monoidal category with duals. Roughly speaking, this amounts to equipping each object with a choice of left adjunction, and requiring the functor $F \colon C_{1,1} \to X$ to preserve this choice. Then $F$ is determined up to monoidal unitary natural transformation by the requirement that it map the positively oriented point to a particular object $x \in X$. In this paper we have attempted to take the `weak' rather than the `strict' approach. Our point here is that the weak approach seems to make it more difficult to formulate the sense in which $C_{1,1}$ is the `free' monoidal category with duals on one object. In higher dimensions the balancing plays an interesting role in this issue. Framed oriented tangles in 3 dimensions form a braided monoidal category $C_{1,2}$ with duals. Here by `braided monoidal category with duals' we mean a monoidal category with duals which is also braided, such that the braiding is unitary and every object $x$ has a well-balanced adjunction $(x,x^\ast,i,e)$. For any object $x$ in a braided monoidal category $X$ with duals, there is a braided monoidal $\ast$-functor $F \colon C_{1,2} \to X$ sending the positively oriented point to $x$. Moreover, because well-balanced adjunctions are unique up to unique {\it unitary} isomorphism, $F$ is unique up to monoidal unitary natural isomorphism. This gives a sense in which $C_{1,2}$ is the free braided monoidal category with duals on one object. Similarly, framed oriented tangles in 4 dimensions form a symmetric monoidal category with duals $C_{1,3}$, i.e., a braided monoidal category with duals for which the braiding is a symmetry. Again, for any object $x$ in a symmetric monoidal category $X$ with duals, there is a symmetric monoidal $\ast$-functor $F \colon C_{1,3} \to X$ sending the positively oriented point to $x$, and $F$ is unique up to monoidal unitary natural isomorphism. (For an alternative `strict' approach to the 3- and 4-dimensional cases, see HDA0.) In short, we need to understand the notion of $k$-tuply monoidal $n$-Hilbert spaces more deeply, as well as the notion of `free' $k$-tuply monoidal $n$-categories with duals. \subsection*{Acknowledgements} Many of the basic ideas behind this paper were developed in collaboration with James Dolan. I would also like to thank Louis Crane, Martin Hyland, Martin Neuchl, John Power, Stephen Sawin, Gavin Wraith, and David Yetter for helpful conversations and correspondence. I am grateful to the Erwin Schr\"odinger Institute and the physics department of Imperial College for their hospitality while part of this work was being done.
proofpile-arXiv_065-477
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} Matter containing a large number of strange quarks may have a lower energy per baryon than ordinary nuclei and be absolutely stable\cite{1,2}. This intriguing possibility has generated a great deal of interest across many subfields of physics. A crucial implicit assumption for the stability of this new kind of matter, called strange matter, is that the baryon number $B$ is exactly conserved. Of course, we know nuclei are stable against $\Delta B \neq 0$ decays. The best mode-independent experimental lower limit to date is $1.6 \times 10^{25}$ years\cite{3}. However, there may be effective $\Delta B \neq 0$ interactions which are highly suppressed for nuclei but not for strange matter. We examine such a hypothesis here and show that whereas such exotic decays are possible, the relevant lifetime should be longer than $10^5$ years. In Sec.~2 we review briefly the present experimental constraints on $\Delta B \neq 0$ interactions. These come mainly from the nonobservation of proton decay and of neutron-antineutron oscillation. We then point out the possible consequences of an effective $\Delta B = 2$, $\Delta N_s = 4$ operator (where $N_s$ denotes the number of $s$ quarks) on the stability of strange matter. We proceed to discuss some possible theoretical origins of such an operator in Sec.~3 and Sec.~4. We deal first with the supersymmetric standard model with $R$-parity nonconserving terms of the form $\lambda_{ijk} u_i^c d_j^c d_k^c$ and show that their effects are negligible. We then discuss two possible extended models where the effects may be large. In Sec.~5 we consider this operator in conjunction with the usual weak interaction and show that the suppression coming from the stability of ordinary nuclei translates to a phenomenological lower limit of $10^5$ years on the lifetime of strange matter. Finally in Sec.~6, there are some concluding remarks. \section{Baryon-Number Nonconserving Interactions} Given the standard $SU(3) \times SU(2) \times U(1)$ gauge symmetry and the usual quarks and leptons, it is well-known that the resulting renormalizable Lagrangian conserves both the baryon number $B$ and the lepton number $L$ automatically. However, new physics at higher energy scales may induce effective interactions (of dimension 5 or above) which do not respect these conservation laws. The foremost example is the possibility of proton decay. For this to happen, there has to be at least one fermion with a mass below that of the proton. Since only leptons are known to have this property, the selection rule $\Delta B = 1$, $\Delta L = 1$ is applicable. The most studied decay mode experimentally is $p \rightarrow \pi^0 e^+$, which requires the effective interaction \begin{equation} {\cal H}_{int} \sim {1 \over M^2} (uude). \end{equation} It is now known\cite{4} that $\tau_{exp} (p \rightarrow \pi^0 e^+) > 9 \times 10^{32}$ years, hence $M > 10^{16}$ GeV. [Here $M$ should be considered an effective mass, because in some scenarios the denominator of the right-hand side of Eq.~(1) is really the product of two different masses.] This means that proton decay probes physics at the grand-unification energy scale. The next simplest class of effective interactions has the selection rule $\Delta B = 2$, $\Delta L = 0$. The most well-known example is of course \begin{equation} {\cal H}_{int} \sim {1 \over M^5} (udd)^2, \end{equation} which induces neutron-antineutron oscillation and allows a nucleus to decay by the annihilation of two of its nucleons. Note that since 6 fermions are now involved, the effective operator is of dimension 9, hence $M$ appears to the power $-5$. This means that for the same level of nuclear stability, the lower bound on $M$ will be only of order $10^5$ GeV. The present best experimental lower limit on the $n - \bar n$ oscillation lifetime is\cite{5} $8.6 \times 10^7$s. Direct search for $N N \rightarrow \pi \pi$ decay in iron yields\cite{6} a limit of $6.8 \times 10^{30}$y. Although the above two numbers differ by 30 orders of magnitude, it is well established by general arguments as well as detailed nuclear model calculations that\cite{7} \begin{equation} \tau (n \bar n) = 8.6 \times 10^7{\rm s} ~ \Rightarrow ~ \tau (N N \rightarrow \pi \pi) \sim 2 \times 10^{31}{\rm y}. \end{equation} Hence the two limits are comparable. To estimate the magnitude of $M$, we use \begin{equation} \tau(n \bar n) = M^5 |\psi(0)|^{-4}, \end{equation} where the effective wavefunction at the origin is roughly given by\cite{8} \begin{equation} |\psi(0)|^{-2} \sim \pi R^3, ~~~ R \sim 1 ~ {\rm fm}. \end{equation} Hence we obtain $M > 2.4 \times 10^5$ GeV. Bearing in mind that ${\cal H}_{int}$ is likely to be suppressed also by products of couplings less than unity, this means that the stability of nuclei is sensitive to new physics at an energy scale not too far above the electroweak scale of $10^2$ GeV. It may thus have the hope of future direct experimental exploration. Consider next the effective interaction \begin{equation} {\cal H}_{int} \sim {1 \over M^5} (uds)^2. \end{equation} This has the selection rule $\Delta B = 2$, $\Delta N_s = 2 ~(\Delta S = -2)$. In this case, a nucleus may decay by the process $N N \rightarrow K K$\cite{9,10}. Although such decay modes have never been observed, the mode-independent stability lifetime of $1.6 \times 10^{25}$y mentioned already is enough to guarantee that it is very small. However, consider now \begin{equation} {\cal H}_{int} \sim {1 \over M^5} (uss)^2, \end{equation} which has the selection rule \begin{equation} \Delta B = 2, ~~~ \Delta N_s = 4. \end{equation} To get rid of four units of strangeness, the nucleus must now convert two nucleons into four kaons, but that is kinematically impossible. Hence the severe constraint from the stability of nuclei does not seem to apply here, and the following interesting possibility may occur. Strangelets ({\it i.e.} stable and metastable configurations of quarks with large strangeness content) with atomic number $A$ (which is of course the same as baryon number $B$) and number of strange quarks $N_s$ may now decay into other strangelets with two less units of $A$ and one to three less $s$ quarks. For example, \begin{eqnarray} (A, N_s) &\rightarrow& (A - 2, N_s - 2) + K K, \\ (A, N_s) &\rightarrow& (A - 2, N_s - 1) + K K K, ~etc. \end{eqnarray} For the states of lowest energy, model calculations show\cite{11} that $N_s \sim 0.8 A$, hence the above decay modes are efficient ways of reducing all would-be stable strangelets to those of the smallest $A$. Unlike nuclei which are most stable for $A$ near that of iron, the energy per baryon number of strangelets decreases with increasing $A$. This has led to the intriguing speculation that there are stable macroscopic lumps of strange matter in the Universe. On the other hand, it is not clear how such matter would form, because there are no stable building blocks such as hydrogen and helium which are essential for the formation of heavy nuclei. In any case, the above exotic interaction would allow strange matter to dissipate into smaller and smaller units, until $A$ becomes too small for the strangelet itself to be stable. A sample calculation by Madsen\cite{11} shows that \begin{equation} m_0 (A) - m_0 (A - 2) \simeq (1704 + 111 A^{-1/3} + 161 A^{-2/3})~{\rm MeV}, \end{equation} assuming $m_s = 100$ MeV, and the bag factor $B^{1/4} = 145$ MeV. Hence the $KK$ and $KKK$ decays would continue until the ground-state mass $m_0$ exceeds the condition that the energy per baryon number is less than 930 MeV, which happens at around $A = 13$. \section{Supersymmetric Standard Model} If the standard model of quarks and leptons is extended to include supersymmetry, the $\Delta B \neq 0$ terms $\lambda_{ijk} u_i^c d_j^c d_k^c$ are allowed in the superpotential. In the above notation, all chiral superfields are assumed to be left-handed, hence $q^c$ denotes the left-handed charge-conjugated quark, or equivalently the right-handed quark, and the subscripts refer to families, {\it i.e.} $u_i$ for $(u, c, t)$ and $d_j$ for $(d, s, b)$. Since all quarks are color triplets and the interaction must be a singlet which is antisymmetric in color, the two $d$ quark superfields must belong to different families. In the minimal supersymmetric standard model, these terms are forbidden by the imposition of $R$-parity. However, this assumption is not mandatory and there is a vast body of recent literature exploring the consequences of $R$-parity nonconservation\cite{12}. Starting with Yukawa terms of the form $\lambda_{ijk} u_i^c d_j^c d_k^c$, where two of the fields are quarks and the third is a scalar quark, we can obtain the effective interaction of Eq.~(7) in several ways\cite{13}. In Ref.~[10] the $d d \rightarrow \tilde b \tilde b$ box diagram (where $\tilde q$ denotes the supersymmetric scalar partner of $q$) is considered to obtain the effective interaction of Eq.~(2) for $n - \bar n$ oscillation using $\lambda_{udb}$. Here we take \begin{equation} s s \rightarrow \tilde b \tilde b \end{equation} and use $\lambda_{usb}$ instead. The box diagram involves the exchange of the $(u,c,t)$ quarks, the $W$ boson, and their supersymmetric partners. Since only gauge couplings appear, the Glashow-Iliopoulos-Maiani (GIM) mechanism\cite{14} is operative and this diagram vanishes if the $(\tilde u,\tilde c,\tilde t)$ scalar quarks have the same mass and that the $(u,c,t)$ quark masses can be neglected. Thus in Fig.~3 of Ref.~[10], there is a peak at $M_{\tilde t} = 200$ GeV. However, there are actually two scalar quarks $(\tilde q_L,\tilde q_R)$ for each quark $q$. Consider just $t$ and $(\tilde t_L,\tilde t_R)$. The $\tilde t$ mass matrix is given by \begin{equation} {\cal M}^2_{\tilde t} = \left[ \begin{array} {c@{\quad}c} \tilde m_L^2 + m_t^2 & A m_t \\ A m_t & \tilde m_R^2 + m_t^2 \end{array} \right]. \end{equation} Since only $\tilde t_L$ couples to $b_L$ through the gauge fermion $\tilde w$ and $\tilde t_L$ is not a mass eigenstate, its approximate effective contribution is given by \begin{equation} M_{\tilde t_L}^{-2} = {{\tilde m_R^2 + m_t^2} \over {(\tilde m_L^2 + m_t^2) (\tilde m_R^2 + m_t^2) - A^2 m_t^2}}. \end{equation} Unless this somehow cancels the $\tilde u$ and $\tilde c$ contributions accidentally, the box diagram will not vanish. Furthermore, it is often assumed that the soft supersymmetry-breaking terms $\tilde m_L^2$ and $\tilde m_R^2$ are universal, in which case this amplitude would be zero if the $(u,c,t)$ quark masses were neglected. Rewriting Eq.~(7) more specifically as \begin{equation} {\cal H}_{int} = T_1 \epsilon_{\alpha \beta \gamma} \epsilon_{\alpha' \beta' \gamma'} u_{R \alpha} s_{R \beta} s_{L \gamma} u_{R \alpha'} s_{R \beta'} s_{L \gamma'}, \end{equation} where the color indices and chiralities of the quarks are noted, the contribution of Eq.~(12) is then given by\cite{15} \begin{equation} T_1 \simeq {{g^4 \lambda^2_{usb} A^2 m_b^2 m_{\tilde w}} \over {32 \pi^2 \tilde m^8}} V_{ts}^2 F(m^2_{\tilde w}, M_W^2, \tilde m^2, m_t, A), \end{equation} where we have assumed universal soft supersymmetry-breaking scalar masses, $V_{ts}$ is the quark-mixing entry for $t$ to $s$ through the $W$ boson, and \begin{eqnarray} F &=& {1 \over 2} J(m^2_{\tilde w}, M_W^2, \tilde m^2 + m_t^2 + A m_t, m_t^2) + {1 \over 2} J(m^2_{\tilde w}, M_W^2, \tilde m^2 + m_t^2 - A m_t, m_t^2) \nonumber \\ &-& J(m^2_{\tilde w}, M_W^2, \tilde m^2, m_t^2) - \{ m_t^2 \rightarrow 0 ~{\rm in~the~last~entry~of~each~term} \}, \end{eqnarray} with the function $J$ given by\cite{10,13} \begin{equation} J(a_1, a_2, a_3, a_4) = \sum_{i=1}^4 {{a_i^2 \ln (a_i)} \over {\prod_{k \neq i} (a_i - a_k)}}. \end{equation} Using the correspondence of $n - \bar n$ oscillation to $N N$ annihilation inside a nucleus, we estimate the lifetime of strangelets from the above effective interaction assuming $A = 200$ GeV, $\lambda_{usb} < 1$, and $\tilde m > 200$ GeV, to be \begin{equation} \tau > 10^{22}{\rm y}. \end{equation} This tells us that such contributions are negligible from the supersymmetric standard model. Recently, another contribution to Eq.~(2) has been identified\cite{13} involving $\lambda_{tds}$ and $\lambda_{tdb}$ without the GIM suppression of Eq.~(16). The form of its contribution to Eq.~(7) is \begin{equation} {\cal H}_{int} = T_2 \epsilon_{\alpha \beta \gamma} \epsilon_{\alpha' \beta' \gamma'} u_{L \alpha} s_{L \beta} s_{R \gamma} u_{L \alpha'} s_{L \beta'} s_{R \gamma'}, \end{equation} which is not the same as Eq.~(15). The important coupling is now $\lambda_{tsb}$, but $T_2$ is also suppressed relative to $T_1$ by $V_{ub}^2$, hence this contribution to Eq.~(7) is even more negligible. \section{Extended Models Involving Strangeness} Although the supersymmetric standard model cannot have a sizable contribution to the $\Delta B = 2$, $\Delta N_s = 4$ operator of Eq.~(7), an extended model including additional particles belonging to the fundamental {\bf 27} representation of $E_6$, inspired by superstring theory\cite{16}, may do better. Consider a slight variation of the model of exotic baryon-number nonconservation proposed some years ago\cite{17}. In addition to the usual quark and lepton superfields \begin{equation} Q \sim (3, 2, 1/6), ~~~ u^c \sim (\bar 3, 1, -2/3), ~~~ d^c \sim (\bar 3, 1, 1/3); \end{equation} \begin{equation} L \sim (1, 2, -1/2), ~~~ e^c \sim (1, 1, 1), ~~~ N^c \sim (1, 1, 0); \end{equation} we have \begin{equation} h \sim (3, 1, -1/3), ~~~ h^c \sim (\bar 3, 1, 1/3); \end{equation} \begin{equation} E \sim (1, 2, -1/2), ~~~ \bar E \sim (1, 2, 1/2), ~~~ S \sim (1, 1, 0); \end{equation} each transforming under the standard $SU(3) \times SU(2) \times U(1)$ gauge group as indicated. Imposing the discrete symmetry $Z_2 \times Z_2$ under which \begin{eqnarray} Q, u^c, d^c, N^c &\sim& (+, -), \\ L, e^c &\sim& (-, +), \\ h, h^c, E, \bar E, S &\sim& (+, +), \end{eqnarray} we get the allowed terms \begin{equation} Q Q h, ~~ u^c d^c h^c, ~~ d^c h N^c, ~~ N^c N^c, \end{equation} in the superpotential, but not $d^c h$. We also assume that $\tilde N^c$ does not develop a vacuum expectation value, so that $B$ is broken explicitly by the above $N^c N^c$ term only. [Note that without this last term, we can assign $B = -2/3, 2/3, 1$ to $h, h^c, N^c$ respectively and $B$ would be conserved.] This model allows us to obtain an effective $(uss)^2$ operator without going through a loop, as shown in Fig.~1. Using the formalism of Ref.~[10] for the process \begin{equation} (A, N_s) \rightarrow (A-2, N_s-2) + KK, \end{equation} we estimate the lifetime to be \begin{equation} \tau \sim {{32 \pi m_N^2} \over {9 \rho_N}} \left[ {M \over \tilde \Lambda} \right]^{10} \sim 1.2 \times 10^{-28} \left[ {M \over \tilde \Lambda} \right]^{10} {\rm y}, \end{equation} where $m_N \sim 1$ GeV, $\rho_N \sim 0.25$ fm$^{-3}$ is the nuclear density, and $\tilde \Lambda \sim 0.3$ GeV is the effective interaction energy scale corresponding to Eq.~(5). If we assume $M = 1$ TeV, then \begin{equation} \tau \sim 2 \times 10^7 {\rm y}. \end{equation} In the above, we have assumed that $\lambda_{ush}$ is unconstrained phenomenologically. However, consider the term $Q_1 Q_2 h$ where $Q_1 = (u, d')$ and $Q_2 = (c, s')$ so that $Q_1 Q_2 = u s' - c d'$. Hence from this term alone, $\lambda_{udh} = (V_{cd}/V_{cs}) \lambda_{ush}$. Since $\lambda_{udh}$ may now combine with $\lambda_{s^c h N^c}$ to induce $N N \rightarrow K K$ decays, its magnitude is seriously suppressed and $\lambda_{ush}$ would be too small for Eq.~(31) to be valid. However, there is also a third generation, allowing us the $Q_1 Q_3 = u b' - t d'$ term which may then be fine-tuned to eliminate the $\lambda_{udh}$ component, thus saving Eq.~(31). This is of course not a very natural solution, but cannot otherwise be ruled out. We may also consider the following tailor-made extension. Let there be a new exotic scalar multiplet $\tilde Q_6 \sim (\bar 6, 1, 2/3)$ with $B = -2/3$ and a new exotic fermion multiplet $\psi_8 \sim (8, 1, 0)$ with $B = 1$. Assume the existence of Yukawa terms $s^c s^c \tilde Q_6^*$, $u^c \psi_8 \tilde Q_6$, and the $B$-nonconserving Majorana mass term $\psi_8 \psi_8$, then an effective $(uss)^2$ term is possible, as shown in Fig.~2. Note that the $uss$ combination is now a color octet. This may in fact be a more efficient way to dissipate strange matter which is presumably not clumped into color-singlet constituents as in ordinary nuclei. \section{Stability Limit of Strange Matter} Since $\tau$ depends on $M/\tilde \Lambda$ to the power 10 in Eq.~(30), it appears that a much shorter lifetime than that of Eq.~(31) is theoretically possible. However, there is a crucial phenomenological constraint which we have yet to consider. Although two nucleons cannot annihilate inside a nucleus to produce four kaons, they can make three kaons plus a pion. The effective $(uss)^2$ operator must now be supplemented by a weak transition $s \rightarrow u + d + \bar u$. We can compare the effect of this on $N N \rightarrow K K K \pi$ versus that of the $(uds)^2$ operator on $N N \rightarrow K K$ discussed in Ref.~[10]. First let us look at the phase-space difference. Consider a nucleus with atomic number $A$ decaying into one with atomic number $A-2$ plus two kaons with energy-momentum conservation given by $p = p' + k_1 + k_2$. The decay rate is proportional to \begin{eqnarray} F_1 &=& {1 \over {(2 \pi)^5}} \int {{d^3 k_1} \over {2 E_1}} \int {{d^3 k_2} \over {2 E_2}} \int {{d^3 p'} \over {2 E'}} \delta^4 (p - p' - k_1 - k_2) \nonumber \\ &\simeq& {1 \over {2 M'}} {1 \over {(2 \pi)^5}} \int {{d^3 k_1} \over {2 E_1}} \int {{d^3 k_2} \over {2 E_2}} \delta (M - M' - E_1 - E_2) \nonumber \\ &\simeq& {1 \over {2 M'}} {1 \over {(2 \pi)^3}} \int_{m_K}^{E_{max}} k_1 k_2 dE_1, \end{eqnarray} where $k_1 = (E_1^2 - m_K^2)^{1/2}$, $k_2 = [(2 m_N - E_1)^2 - m_K^2]^{1/2}$, and $E_{max} = 2 m_N - m_K$. \newpage Consider next $A \rightarrow (A-2) + K K K \pi$ with $p = p' + k_1 + k_2 + k_3 + k_\pi$. Because of the limited phase space, we assume the kaons to be nonrelativistic. Hence \begin{eqnarray} F_2 &\simeq& {1 \over {2 M'}} {1 \over {(2 \pi)^{11}}} \int {{d^3 k_1} \over {2 m_K}} \int {{d^3 k_2} \over {2 m_K}} \int {{d^3 k_3} \over {2 m_K}} \int {{d^3 k_\pi} \over {2 E_\pi}} \delta (M - M'- 3 m_K - {{k_1^2 + k_2^2 + k_3^2} \over {2 m_K}} - E_\pi) \nonumber \\ &\simeq& {1 \over {2 M'}} {1 \over {(2 \pi)^7}} {1 \over m_K^3} \int k_\pi k_1^2 k_2^2 k_3^2 dk_1 dk_2 dk_3, \end{eqnarray} where $k_\pi = \{[2m_N - 3m_K - (k_1^2 + k_2^2 + k_3^2)/2m_k]^2 - m_\pi^2 \}^{1/2}$. The above integral can be evaluated by treating $k_{1,2,3}$ as Cartesian coordinates and then convert them to three-dimensional polar coordinates. The angular integration over the $k_{1,2,3} > 0$ octant yields a factor of $\pi/210$ and we get \begin{equation} F_2 \simeq {1 \over {2 M'}} {1 \over {(2 \pi)^7}} {1 \over m_K^3} {\pi \over {210}} \int_0^{k_{max}} k_\pi k^8 dk, \end{equation} where $k_{max} = [2 m_K (2 m_N - 3 m_K - m_\pi)]^{1/2}$. The effective interaction here also differs from that of Eq.~(32) by the appearance of a third kaon and an extra pion which can be thought of as having been converted by the weak interaction from a fourth kaon. We thus estimate the suppression factor to be \begin{equation} \xi \sim \left[ {{f_K^2 T_{K \pi}} \over {\tilde \Lambda^6}} \right]^2 {F_2 \over F_1} \sim 10^{-20}, \end{equation} where $f_K = 160$ MeV, and $T_{K \pi} = 0.07$ MeV$^2$ is obtained using the symmetric soft-pion reduction\cite{18} of the experimental $K \rightarrow 2 \pi$ amplitude. We have again invoked the effective interaction energy scale $\tilde \Lambda = 0.3$ GeV used in Eq.~(30). Obviously, our estimate depends very sensitively on this parameter, but that is not untypical of many calculations in nuclear physics. Since the stability of nuclei is at least $1.6 \times 10^{25}$ years, the reduction by $\xi$ of the above yields \begin{equation} \tau > 10^5 {\rm y} \end{equation} for the lifetime of strangelets against $\Delta B = 2$, $\Delta N_s = 4$ decays. \section{Concluding Remarks} We have pointed out in this paper that an effective $(uss)^2$ interaction may cause stable strange matter to decay, but the lifetime is constrained phenomenologically by the stability of nuclei against decays of the type $A \rightarrow (A-2) + K K K \pi$ and we estimate it to have a lower limit of $10^5$ years. This result has no bearing on whether stable or metastable strangelets can be created and observed in the laboratory, but may be important for understanding whether there is stable strange bulk matter left in the Universe after the Big Bang and how it should be searched for. For example, instead of the usual radioactivity of unstable nuclei, strange matter may be long-lived kaon and pion emitters. The propagation of these kaons and pions through the matter itself and their interactions within such an environment are further areas of possible study. Since energy is released in each such decay, although the amount is rather small, it may be sufficient to cause a chain reaction and break up the bulk matter in a cosmological time scale. Furthermore, if the particles mediating this effective interaction have masses of order 1 TeV as discussed, then forthcoming future high-energy accelerators such as the Large Hadron Collider (LHC) at the European Center for Nuclear Research (CERN) will have a chance of confirming or refuting their existence. \vspace{0.3in} \begin{center} {ACKNOWLEDGEMENT} \end{center} This work was supported in part by the U. S. Department of Energy under Grant No. DE-FG03-94ER40837. \newpage \bibliographystyle{unsrt}
proofpile-arXiv_065-478
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section*{Note Added:} \noindent After this letter was accepted for publication, we became aware of the work of Girvin and MacDonald \cite{girvin}, where they showed that the gauge-transformed Laughlin wave-function [ eq. (7) of their paper] shows off-diagonal long-range order. It then immediately follows that the Calogero-Sutherland ground state wave-function in two dimensions as given by \eq{grst} [which is identical to eq. (7) of \cite{girvin}] also exhibits off-diagonal long-range order.
proofpile-arXiv_065-479
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section*{Introduction} Random sequential adsorption (RSA) \cite{Evans-93} is an irreversible process which particles are deposited randomly and consecutively on a surface. The depositing particles, represented by hard-core extended objects, satisfy the excluded volume condition where they are not allowed to overlap. The exclusion of certain regions for further deposition attempts due to the adsorbed particles leads to a dominant infinite-memory correlation effect where the system approaches partially covered, fully blocked stage at large times. However, this picture is altered when the diffusional relaxation is introduced \cite{Privman-Nielaba-92,Nielaba-Privman-92,% Wang-group-93}. Privman and Nielaba \cite{Privman-Nielaba-92} have shown that the effect of added diffusional relaxation in the deposition of dimer on a 1D lattice substrate is to allow the full, saturation coverage via a $\sim t^{-1/2}$ power law at large times, preceded by a mean-field crossover regime with the intermediate $\sim t^{-1}$ behavior for fast diffusion. Series expansion is one of the powerful analytical methods in the RSA studies \cite{Baram-Kutasov-89,Dickman-Wang-Jensen-91,% Oliveira-group-92,Bonnier-group-93,% Baram-Fixman-95,Gan-Wang-96}. Long series in powers of time $t$ have been obtained, reminiscent of series expansions in equilibrium statistical mechanics, by using a computer \cite{Martin-74}. Recently, the authors \cite{Gan-Wang-96} have proposed an efficient algorithm for generating long series for the coverage $\theta$ in powers of time $t$ based on the hierarchical rate equations. The present work is to study the time-dependent quantity $\theta$ for one-dimensional models of RSA with diffusional relaxation, both analytically and numerically. It will be seen that even though relatively long series have been obtained, we are still unable to extract the kinetics of the systems at large times for general $\gamma$ due to long, rich transient crossover regime that the series must describe. Extensive computer simulations are performed to confirm the $t^{-1/2}$ power law approach of $\theta$, where we have employed an efficient event-driven algorithm. The remainder of this paper is organized as follows. Section~\ref{sec:model} introduces two related models. Details of series expansion are explained in Section~\ref{sec:series_expansion}. Analyses of the series can be found in Section~\ref{sec:series_analysis}. Monte Carlo results are presented in Section~\ref{sec:monte_carlo} and finally Section~\ref{sec:conclusion} contains the summary and conclusions. \section{The models} \label{sec:model} Two models have been studied in this work. We start with an initially empty, infinite linear lattice. Dimers are dropped randomly and sequentially at a rate of $k$ per lattice site per unit time, onto the lattice. Hereafter we set $k$ equal to unity without loss of generality. If the chosen two neighbor sites are unoccupied, the dimer is adsorbed on the lattice. If one of the chosen sites is occupied, the adsorption attempt is rejected. One of the simplest possibilities of diffusional relaxation in this dimer adsorption process is that the adsorbed dimer is permitted to hop either to left or right by one lattice constant at a diffusion rate $\gamma$ from the original dimer position, provided that the diffusion attempt does not violate the excluded volume condition. This model has been initiated and studied by Privman and Nielaba \cite{Privman-Nielaba-92}. We refer this model as the dimer RSA with dimer diffusion or diffusive dimer model. A second possibility is that an adsorbed dimer is allowed to dissociate into two independent monomers; each monomer can diffuse to one of its nearest neighbor sites with a diffusion rate $\gamma$, provided that the diffusion attempt does not violate the excluded volume condition. This model bears a strong resemblance to the former model and is exactly solvable when $\gamma = 1/2$ \cite{Grynberg-Stinchcombe-95}. We refer this model as the dimer RSA with monomer diffusion or diffusive monomer model. Interestingly enough, the special case of the diffusive monomer problem with $\gamma = 1/2$ can be mapped to the diffusion-limited process \begin{equation} {\cal A} + {\cal A} \rightarrow \hbox{inert}, \end{equation} which is known as one-species annihilation process \cite{Avraham-group-90}. This model has been solved exactly by a number of researchers \cite{Lushnikov-87,Spouge-88,Balding-group-88}. We observe that when $\gamma = 1/2$, the effect of a dimer deposition attempt in the diffusive monomer model corresponds to two diffusion attempts of $\cal A$ in an adjacent pair of $\cal A$ of the ${\cal A} + {\cal A} \rightarrow \hbox{inert}$ process. The time-dependent quantity coverage $\theta(t)$ (fraction of occupied sites) for the diffusive monomer model with $\gamma = 1/2$ is given by \begin{eqnarray} \theta(t) = 1- \exp(-2t)I_0(2t), \end{eqnarray} where $I_n(z) $ is the modified Bessel function of integer order $n$. \unitlength=6pt \def\circle{0.6}{\circle{0.6}} \def\circle*{0.6}{\circle*{0.6}} \def\o{\begin{picture}(1,1)(-0.5,-0.5) \put(0,0){\circle{0.6}} \end{picture}} \def\x{\begin{picture}(1,1)(-0.5,-0.5) \put(0,0){\circle*{0.6}} \end{picture}} \def\oo{\begin{picture}(2,1)(-0.5,-0.5) \put(0,0){\circle{0.6}} \put(1,0){\circle{0.6}} \end{picture}} \def\ooo{\begin{picture}(3,1)(-0.5,-0.5) \put(0,0){\circle{0.6}} \put(1,0){\circle{0.6}} \put(2,0){\circle{0.6}} \end{picture}} \def\oox{\begin{picture}(3,1)(-0.5,-0.5) \put(0,0){\circle{0.6}} \put(1,0){\circle{0.6}} \put(2,0){\circle*{0.6}} \end{picture}} \def\oxo{\begin{picture}(3,1)(-0.5,-0.5) \put(0,0){\circle{0.6}} \put(1,0){\circle*{0.6}} \put(2,0){\circle{0.6}} \end{picture}} \def\oooo{\begin{picture}(4,1)(-0.5,-0.5) \put(0,0){\circle{0.6}} \put(1,0){\circle{0.6}} \put(2,0){\circle{0.6}} \put(3,0){\circle{0.6}} \end{picture}} \def\ooox{\begin{picture}(4,1)(-0.5,-0.5) \put(0,0){\circle{0.6}} \put(1,0){\circle{0.6}} \put(2,0){\circle{0.6}} \put(3,0){\circle*{0.6}} \end{picture}} \def\ooxo{\begin{picture}(4,1)(-0.5,-0.5) \put(0,0){\circle{0.6}} \put(1,0){\circle{0.6}} \put(2,0){\circle*{0.6}} \put(3,0){\circle{0.6}} \end{picture}} \def\ooxx{\begin{picture}(4,1)(-0.5,-0.5) \put(0,0){\circle{0.6}} \put(1,0){\circle{0.6}} \put(2,0){\circle*{0.6}} \put(3,0){\circle*{0.6}} \end{picture}} \def\oxxo{\begin{picture}(4,1)(-0.5,-0.5) \put(0,0){\circle{0.6}} \put(1,0){\circle*{0.6}} \put(2,0){\circle*{0.6}} \put(3,0){\circle{0.6}} \end{picture}} \def\oooxx{\begin{picture}(5,1)(-0.5,-0.5) \put(0,0){\circle{0.6}} \put(1,0){\circle{0.6}} \put(2,0){\circle{0.6}} \put(3,0){\circle*{0.6}} \put(4,0){\circle*{0.6}} \end{picture}} \def\ooxxo{\begin{picture}(5,1)(-0.5,-0.5) \put(0,0){\circle{0.6}} \put(1,0){\circle{0.6}} \put(2,0){\circle*{0.6}} \put(3,0){\circle*{0.6}} \put(4,0){\circle{0.6}} \end{picture}} \section{Series expansions} \label{sec:series_expansion} To illustrate how series expansions are performed, we note that the first few rate equations for the dimer and monomer diffusive models are \begin{eqnarray} {dP(\o) \over dt} & = & -2 P(\oo), \label{eq:1stdimer}\\ {dP(\oo) \over dt} & = & -P(\oo) - 2P(\ooo) - 2 \gamma P(\ooxx) + 2 \gamma P(\oxxo), \label{eq:2nddimer} \\ {dP(\ooo)\over dt} & = & -2P(\ooo) - 2P(\oooo) - 2\gamma P(\oooxx) +2 \gamma P(\ooxxo), \\ \cdots \nonumber \end{eqnarray} and \begin{eqnarray} {dP(\o) \over dt} & = & -2 P(\oo), \\ {dP(\oo) \over dt} & = & -P(\oo) - 2P(\ooo) - 2\gamma P(\oox) + 2\gamma P(\oxo), \\ {dP(\ooo) \over dt} & = & -2P(\ooo) - 2P(\oooo) - 2\gamma P(\ooox) + 2\gamma P(\ooxo), \\ {dP(\oox)\over dt} & = & -P(\oox) + P(\oooo) - \gamma P(\oox) + \gamma P(\oxo) - \nonumber \\ & & \gamma P(\ooxo) + \gamma P(\ooox), \\ \cdots \nonumber \end{eqnarray} respectively, where $P(C)$ denotes the probability of finding a configuration $C$ of sites specified empty `$\o$' or filled `$\x$'. Unspecified sites can be occupied or empty. Here we have taken into account the symmetries of a configuration under all lattice group operations. For the one-dimensional configurations, we just need to consider the reflection operation only. Let $C_o$ denote a particular configuration of interest, and $P_{{C_o}} \equiv P(C_o)$ the associated configuration probability. $P_{C_o}$ is expected to be a well behaved function of time $t$, so one can obtain the Taylor series expansion with the expansion point at $t = 0$, $P_{C_o}(t) = \sum\limits_{n=0}^{\infty} {P_{C_o}}^{(n)}t^n/n! $, with the $n$th {\it derivative} of $P_{C_o}$ given by \begin{equation} {P_{C_0}}^{(n)} = \left. {d^n P_{C_0}(t) \over d t^n}\right|_{t=0}. \end{equation} Let $G_i$ denote the set of new configurations generated in the calculation of the $i$th derivative of $P_{C_0}$, and $G_i^j$ the corresponding $j$th derivatives of the set of configurations. We observe that $G_0^{n-1}$, $G_1^{n-2}$, $\ldots$, $G_{n-1}^0$ (determined at the $(n-1)$th derivative), $G_0^{n-2}$, $G_1^{n-3}$, $\ldots$, $G_{n-2}^0$ (determined at the $(n-2)$th derivative), $\ldots$, $G_0^0$ are predetermined before calculating the $n$th derivative of $P_{C_o}$. In the calculation of $n$th derivative of $P_{C_o}$, we determine systematically $G_0^n$, $G_1^{n-1}$, $\ldots$, $G_{n-1}^1$, $G_{n}^0$, by recursive use of rate equations. This algorithm is efficient since each value in $G_i^{n-i}$, $ 0 \le i \le n$ and the rate equation for a configuration $C$ is generated once only. However, this algorithm consumes the memory quickly as a result of storage of intermediate results. The computation of the expansion coefficients makes use of the isomorphism between a lattice configuration and its binary representation if we map an occupied (empty) site to 1 (0). The data structures used to represent Eq.~(\ref{eq:1stdimer}) and Eq.~(\ref{eq:2nddimer}) are depicted in Fig.~\ref{fig:flow}. A node for a configuration $C$ is characterized by its four components; (i) the representation of $C$ in the computer, (ii) a pointer to the derivatives of $P_C$, ${P_C}^{(n)}$ for $n = 1, 2, 3, \ldots$, (iii) the highest order of derivative $h$ of $P_C$ obtained so far, and (iv) a pointer to a linked list of nodes of configurations (`children') which appear in the right hand side of the rate equation for $P_C$. The linked list contains the associated coefficients for each `child'. The variable $h$ is used so that we know the values of ${P_C}^{(n)}$ where $ 1 \le n \le h$ have already been calculated and can be retrieved when needed. All pointers to the configuration nodes generated during the enumeration process are stored in a hash table or a binary tree to allow efficient checking of the existence of any configuration. Use of the algorithm and data structures allows us to obtain coefficients up to $t^{31}$ and $t^{27}$ (presented in Appendix \ref{appendix:series_coefficient}) for $P(\o,t)$ of the diffusive dimer and monomer models, respectively. \section{Analyses of series} \label{sec:series_analysis} Analytically, we are interested in confirming the power law approach of $t^{-1/2}$ of the coverage $\theta$ at large times for both diffusive dimer and monomer models through the unbiased and biased analyses of the series. The unbiased analysis does not fix the saturation coverage of the system, while the biased analysis assumes the saturation coverage to be the value 1. For the unbiased analysis, let $\theta(t) = 1 - P(\o, t)$ be the time dependent coverage. Let assume that at very large times $t$ the coverage $\theta$ satisfies the equation \begin{eqnarray} \theta(t) = \theta_c - \frac{A(t)}{t^{\delta}}, \label{eq:power_law} \end{eqnarray} where $\theta_c$ is the saturation coverage. $A(t)$ is assumed to be a function of $t$, which tends to a constant value as $t\rightarrow\infty$, and $\delta$ is the exponent that characterizes the saturation approach (we expect to obtain $\delta = 1/2$ from the analysis of the series for all $\gamma$ values). Writing $t = A(t)^{1/\delta}(\theta_c - \theta)^{-1/\delta}$, we see that if we perform a DLog Pad\'e \cite{Baker-61,Hunter-Baker-73,Baker-Hunter-73} analysis to the inverted series $t= t(\theta)$, where \begin{eqnarray} \frac{d}{d\theta} \log t(\theta) = \frac{1}{\delta}\frac{d}{d\theta} \log A(t) - \frac{1}{\delta}\cdot\frac{1}{\theta-\theta_c}, \end{eqnarray} then the power law of Eq.~(\ref{eq:power_law}) implies a simple, isolated pole of $\theta_c$ with an associated residue of $-1/\delta$. Fig.~\ref{fig:explode} shows the plot of the inverted series $t$ versus $\theta$ for the 28-term series with $\gamma = 1/2$ for the monomer diffusive model. For the diffusive dimer problem, the closest real pole to the value 1 (the expected saturation coverage) for [16,15], [15,16], [15,15], [16,14], and [14,16] Pad\'e approximants are shown in Fig.~\ref{fig:dimer_theta}, with the corresponding saturation exponents $\delta$ displayed in Fig.~\ref{fig:dimer_delta}. Similarly we form [14,13], [13,14], [13,13], [14,12], [12,14] Pad\'e approximants for the diffusive monomer problem, where the results are displayed in Fig.~\ref{fig:monomer_theta} and Fig.~\ref{fig:monomer_delta}. Comparing the graphs for these two models, the diffusive dimer series give a better convergence of $\theta_c$ and $\delta$ against $\gamma$ than that for the diffusive monomer series generally, presumably due to the fact that the coefficients of the series of $P(\o, t)$ alternate in signs in the former model. For small $\gamma$ values ($\gamma < 5$), the estimates for $\theta_c$ and $\delta$ are unstable --- different Pad\'e approximants do not agree with one another. The series with $\gamma = 0$ describes a pure lattice RSA behavior \cite{Dickman-Wang-Jensen-91}, where the system approaches the jamming coverage exponentially. Hence we expect the confirmation for power law of Eq.~(\ref{eq:power_law}) is interfered by the exponential behavior of the series when $\gamma$ is small. For $\gamma > 5$, there are physically favorable estimates for $\theta_c$ and $\delta $ where $\theta_c = 1.00 \pm 0.05$ and $\delta = 1.0\pm 0.1$ for $10 < \gamma < 20$, for both models. These results are the manifestations of the transient regime of $t^{-1}$ approach to saturation. The distribution plot of the poles and zeros in the vicinity of $(1, 0) $ is displayed in Fig.~\ref{fig:zeropole} for the $[14, 13]$ Pad\'e approximant for the 28-term series with $\gamma = 1/2$ for the diffusive monomer model. We see that the real pole closest to $(1,0)$ is not distinguished and isolated from the nearby poles and zeros. This explains the difficulty of unbiased analysis that the intermediate crossover effect masks the power law approach at late stages. We also perform biased analyses for the series. This series analysis have been used by Jensen and Dickman \cite{Jensen-Dickman-93} to extract critical exponents from series in powers of time $t$. We define the $F$-transform of $f(t)$ by \begin{equation} F[f(t)] = t \frac{d}{dt} \ln f. \end{equation} If $f \sim A t^{-\alpha}$ for some constant $A$, then $F(t) \rightarrow \alpha$ as $t \rightarrow \infty$. We consider the exponential transformation \begin{equation} \label{eq:transform} z = \frac{1-e^{-bt}}{b}, \end{equation} which proved to be very useful in the analysis of RSA series \cite{Dickman-Wang-Jensen-91,Gan-Wang-96,Jensen-Dickman-93}. This transformation involves a parameter $b$ which cannot be fixed a priori is then followed by the construction of various orders of Pad\'e approximants to the $z$-series. Crossing region is then searched for in the graphs of $\alpha$ versus $b$, the transformation parameter. To illustrate this biased analysis, we take the saturation coverage $\theta_c$ to be 1 and choose $f(t)$ to be $P(\o,t)= \theta_c - \theta(t)$. Since we expect $P(\o,t) \sim t^{-1/2}$ for large times $t$, specifically we have formed [14,13], [13,14], [13,13], [14,12], [12,14] Pad\'e approximants to the $z$-series for the 28-term series with $\gamma = 1/2$ for the diffusive monomer model. We find that the estimates for $\delta$ is $0.5061(5)$, for $ 0.45 < b < 0.50 $, as we can see from Fig.~\ref{fig:biased_exact}. Thus the exact analytical function of $\exp(-2t)I_0(2t)$ serves as a useful guide of this analysis, where the exponent deviates from the value $1/2$ by only about 1\%. Given a value of $\gamma$, we obtain the corresponding estimates of $\delta$ from the first convergence of all Pad\'e approximants by locating the crossing region. The results of $\delta$ estimates for several values of $\gamma$ are presented in Table~\ref{tab:allexp}. The corresponding uncertainties for $\delta$ which reflect the variation of $\delta$ over a range of $b$ are shown in the same table. For the diffusive dimer model, we have formed [16,15], [15,16], [15,15], [16,14], [14,16], [15,14], and [14,15] Pad\'e approximants to the $z$-series. The corresponding graphs are displayed in Fig.~\ref{fig:allexp}. It is seen that for small values of $\gamma$, we obtain small estimates of $\delta$, while for large $\gamma$, $ \delta \rightarrow 1$, suggesting the approach to the limiting saturation is via a mean-field like result, i.e. the $t^{-1}$ power law. Hence we see that even though the exponential transformation Eq.~(\ref{eq:transform}) works well for the exact series of of diffusive monomer model when $\gamma = 1/2$, its use for general $\gamma$ is not very appropriate. We have also tried the transformation $ z = 1 - (1 + bt)^{-1/2}$ to the series for both diffusive models but the convergence is rather poor. We have tried and used a third method of extracting the saturation exponent $\delta$. If we assume that for large enough times $t$, the saturation coverage $\theta$ assumes a power law \begin{equation} 1 - \theta \propto t^{-\delta}, \end{equation} then we expect a plot of $d\ln(1-\theta)/d\ln(t)$ versus $t$ or $\log_{10}(t)$ should give a plateau of constant $-\delta$ values. By forming [14,13], [13,14], [13,13] Pad\'e approximants to the $d\ln(1-\theta)/d\ln(t)$ of the 28-term series for the diffusive monomer model with $\gamma = 1/2$, we observe from Fig.~\ref{fig:illus} that the agreement between different Pad\'e estimates and the exact solution is excellent for $\log_{10}(t)$ up to around $0.9$. For diffusive dimer problem, three Pad\'e approximants of [16,15], [15,16], and [15,15] are formed. The plots of $d\ln(1-\theta)/d\ln(t)$ versus $\log_{10}(t)$ for the diffusive dimer and monomer models, shown in Fig.~\ref{fig:dimerwing} and Fig.~\ref{fig:monomwing}, respectively, are obtained by taking the average of the 3 different Pad\'e estimates. The graphs end before the difference between at least a pair of Pad\'e estimates is more that 0.001. The last estimates in Fig.~\ref{fig:dimerwing} and Fig.~\ref{fig:monomwing} are taken as the estimates for $\delta$ and they are listed in the last two columns of Table~\ref{tab:allexp}. These estimates for $\delta$ are plotted in the same graph for the $F$-transformed analysis for comparisons (Fig.~\ref{fig:allexp}). It is seen that our last method of extracting the saturation exponents appears to be better than the $F$-transform analysis since it yields almost about the same estimates for $\delta$. It does not involve any transformation which is not known in advance that will yield consistent results \cite{Jensen-Dickman-93}. Looking at the ends of the curves in Fig.~\ref{fig:dimerwing} and Fig.~\ref{fig:monomwing}, we are certain that the power law regime is still not reached since the $\delta$ estimates do not seem to converge to a constant value, except the case when $\gamma = 1/2$ for the diffusive monomer model. From this we know that our estimates for $\delta$ do not describe the true power law approach at large times $t$. Such information cannot be found in the $F$-transform analysis. We note that our last method of analyzing the series is easy to use compared to the $F$-transform analysis. \section{Monte Carlo simulations} \label{sec:monte_carlo} To study the short and large time behaviors of the coverage, we have performed extensive and exhaustive simulations for the diffusive dimer and monomer models. For both models, we take an initially empty linear lattice with $N = 20000$ sites with periodic boundary conditions so that the finite-size effects can be ignored. In each Monte Carlo step, a pair of adjacent sites is chosen randomly. The type of attempted process is then decided: deposition with probability $p$, where $ 0 < p \le 1$, or diffusion with probability $(1-p)$. In the case of the deposition attempts, if any one of the chosen sites is occupied, the deposition attempt is rejected (unsuccessful attempt), else the adsorption attempt is accepted. In the case of diffusion, yet another decision is made either to move right or left, with equal probability. If the selected decision is diffusion to the right, we check the selected pair of sites are occupied and its right nearest neighbor site is unoccupied, then the dimer is moved by one lattice constant to the right. The left-diffusion attempts are treated similarly. In contrast to the diffusive dimer model, the diffusive monomer model allows monomers to move by one lattice constant. We define one time unit interval ($\Delta t = 1$) to be during which a deposition attempt is performed for each lattice site. Thus for $N$-site lattice, one unit time corresponds to $N$ deposition attempts, on average. The diffusion rate $\gamma$ relative to the deposition rate, is then $\gamma = (1-p)/2p$. Straightforward simulation procedure, as described above, encounters a serious drawback in which at late stages, most adsorption and diffusion attempts are rejected. In order to study the behavior of the system at large times, we have used an event-driven algorithm to speed up the dynamics of the simulations \cite{Brosilow-group-91,Wang-94}. Let $q$ be the probability that we can make a successful move, then the probability that the first $(i-1)$th trials is unsuccessful, and the $i$th trial is successful is \begin{eqnarray} P_i = q (1-q)^{i-1}, \hbox{\ \ \ } i = 1, 2, 3, \ldots \label{eq:event_driven} \end{eqnarray} If we restrict all trials to be coming from the successful ones, then two consecutive trials are in fact separated by a random variable $i$ in Eq.~(\ref{eq:event_driven}). This distribution can be generated by \begin{eqnarray} i = \left\lfloor\frac{\ln \xi}{\ln(1-q)}\right\rfloor + 1, \end{eqnarray} where $\xi$ is a uniformly distributed random number between 0 and 1. In employing this method, we have to keep and update an active list of successful moves/attempts, where from its length we can evaluate $q$ at any instance. Simulations are performed on a cluster of fast workstations. Our numerical results are obtained for $\gamma = $ 0.05, 0.10, 0.20, $\ldots$, and 6.40, for $t$ up to $2^{20}$. Each data set is averaged over 500 runs, and the longest run take about 150 CPU hours on a HP712/60. The coverage (fraction of occupied sites), $\theta(t)$, is plotted in Fig.~\ref{fig:dimer_phase} and Fig.~\ref{fig:monom_phase} for the diffusive dimer and monomer models, respectively. We have also performed the simulation at $\gamma = 1/2$ for the diffusive monomer model in order to compare the simulation results with the exact results. It is seen that the agreement between them are so good that actually an overlapping line is observed in Fig.~\ref{fig:monom_phase}. For $\gamma = 0$, we have the exact solution \cite{Dickman-Wang-Jensen-91} \begin{eqnarray} \theta(t) = \frac{1-\exp(-2(1-\exp(-t) ))}{2}. \end{eqnarray} For extremely fast diffusion case, i.e., $\gamma = \infty$, exact results have been obtained, where \begin{eqnarray} t(\theta) = \frac{1}{4}\bigl(\frac{1}{1-\theta} - 1\bigr) - \frac{1}{4} \ln(1-\theta), \label{eq:dimer_fast_diff} \end{eqnarray} for the diffusive dimer model \cite{Privman-Barma-92} and \begin{eqnarray} t(\theta) = \frac{1}{2}\bigl(\frac{1}{1-\theta} - 1\bigr). \label{eq:monom_fast_diff} \end{eqnarray} for diffusive monomer model. The approach of $(\theta_c - \theta) \sim t^{-1} $ at all times is obvious for these extremely fast diffusion models. We have included the lines of slope $-1/2$ to indicate the $t^{-1/2}$ power law clearly. It is seen from Fig.~\ref{fig:dimer_phase} and Fig.~\ref{fig:monom_phase} that for $\gamma \ge 3.20$, the system takes a very long time ($t \approx 10^{4}$) before it can enter the final $t^{-1/2}$ regime. This explain why we have difficulty in extracting the actual power law approach from the primitive expansion in time $t$. To further confirm that the saturation approach indeed follows a power law, we have used a scaling analysis. For large times $t$, let us assume that $(1-\theta)$ has the following scaling form \begin{equation} \label{eq:functional} 1-\theta = (\gamma t)^{-1/2}G(\gamma^b/t) , \end{equation} where $G$ is a scaling function and $b$ is a constant to be determined. Eq. (\ref{eq:functional}) requires that $G(u)$ tends to a constant when $u$ tends to 0 \cite{Privman-private}. This is required because for large times $t$, $(1-\theta) \sim t^{-1/2}$. Let further assume that for large $u$, $G(u) \sim u^z$ for some constant $z$. For extremely large $\gamma$, we have $ (1 -\theta) \sim t^{-1}$ (see Eqs. (\ref{eq:dimer_fast_diff}) and (\ref{eq:monom_fast_diff})) hence $z = 1/2$ and $b = 1$. Writing Eq. (\ref{eq:functional}) as \begin{eqnarray*} 1-\theta & = &(1/\gamma)(\gamma/t)^{1/2}G(\gamma/t) \\ & \equiv & (1/\gamma)F(t/\gamma), \end{eqnarray*} we then make log-log plots of $\gamma (1-\theta)$ versus $t/\gamma$ for the diffusive dimer and monomer models as shown in Figs. \ref{fig:dimerscale} and \ref{fig:monomscale}, respectively. It is seen that the data collapse into single curves at large times $t$. Our numerical data confirm not only the power law decay but also a scaling behavior for $(1-\theta)$. \section{Conclusions} \label{sec:conclusion} By using an efficient algorithm based on the hierarchical rate equations, relatively long series are obtained for two models of 1-d RSA with diffusional relaxation. Analyses of series are performed, but it is seen that even though the series are long, we only manage to extract the behaviors of the systems up to intermediate times only. To study the power law of a system at large times $t$, we see that a series which exhibits a continuous crossover behavior in its short and intermediate times ought to be long enough so that various orders of Pad\'e approximants can still converge in the power law regime. Using these long series, we find that the analysis of series based on the ratio method by Song and Poland \cite{Song-Poland-92} was not useful. Specifically, for the diffusive monomer series when $\gamma = 1/2$ (which corresponds to the A + A $\rightarrow$ 0 process in Song and Poland's work), we obtain the saturation exponent (where they have used the symbol $\nu$) $\delta$ $=$ $2.0$, $1.2$, $0.895$, $0.729$, $0.624$, $0.551$, $0.498$, $0.458$, $0.426$, $0.401$, $0.380$, $0.363$, $0.351$, $\ldots$, which is not seen to be converging towards the expected value of $1/2$. We have also performed extensive computer simulations using an efficient event-driven algorithm, where it allows us to use and simulate a larger system to much larger times $t$ than it was done previously on a supercomputer \cite{Privman-Nielaba-92}. The $t^{-1/2}$ power law approach of $\theta$ to its saturation is confirmed numerically at large times $t$. \section*{Acknowledgement} This work was supported in part by an Academic Research Grant RP950601 of National University of Singapore. Part of the calculations were performed on the facilities of the Computation Center of the Institute of Physical and Chemical Research, Japan. We would like to thank V. Privman for pointing out Ref. \cite{Grynberg-Stinchcombe-95} and useful discussion on how to analyse the series. We also appreciate one of the referees for suggesting us to make a scaling behavior study. \pagebreak \bibliographystyle{plain}
proofpile-arXiv_065-480
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{\bf INTRODUCTION} Binary neutron star systems which are spiraling toward their final coalescence under the dissipative influence of gravitational radiation reaction forces are the primary targets for detection of gravitational waves by interferometric gravitational wave detectors such as LIGO and VIRGO\cite{Abramovici92,Thorne95}. Extracting the gravitational waves from the detector noise and making use of the information encoded in the signals will require a thorough knowledge of the expected waveforms produced by these binaries\cite{Thorne95,Cutler93}. In this paper we explore the effect of the neutron-star equation of state on the orbital evolution and gravitational wave emission of binaries just prior to merging. Specifically, we show that a combination of post-Newtonian (relativistic) effects and Newtonian tidal effects (which depend on the equation of state) conspire to induce a dynamical instability in the orbital motion, which causes the plunge to final coalescence to begin somewhat sooner -- and to proceed somewhat faster -- than it would simply under the influence of the dissipative radiation reaction force. Thus the motion of the bodies during the late stages of binary inspiral depends on the structure of the neutron stars. Consequently, the gravitational waveform emitted during this short portion of the final coalescence will be imprinted with information about the nuclear equation of state. During the final $\sim 10$ minutes or the last $\sim 8000$ orbits of a neutron star binary inspiral, the orbital frequency increases from about 5 Hz on up to a cutoff of a few hundreds to a thousand Hertz (roughly corresponding to the orbital frequency when the final plunge begins). Thus the gravitational wave frequency (twice the orbital frequency for the dominant quadrupole radiation) {\it chirps} through the LIGO detector bandwidth during this period \cite{Abramovici92}. The evolution of the binary in these last few minutes of the inspiral is very sensitive to a number of relativistic effects, such as gravitational-wave tails and spin-orbit coupling (dragging of inertial frames). The gravitational waveform emitted by the binary in this portion of the coalescence, the adiabatic inspiral, is currently being extensively studied \cite{bdiww,biww,ww}. During most of this inspiral phase the neutron stars can be treated as simple point masses because the effects associated with the finite stellar size turn out to be small: (i) The neutron star has too small a viscosity to allow for angular momentum transfer from the orbit to the stellar spin via viscous tidal torque\cite{Bildsten92,Kochanek92}; (ii) The effect of the spin-induced quadrupole is negligible unless the neutron star has rotation rate close to the break-up limit\cite{Bildsten92,LRS94}; (iii) Resonant excitations of neutron star internal modes (which occur at orbital frequencies less than $100$ Hz) produce only a small change in the orbital phase due to the weak coupling between the modes and the tidal potential\cite{Lai94,Reisenegger94,Shibata94}; (iv) The correction to the equation of motion from the (static) tidal interaction is of order $(R_o/r)^5$ (where $R_o$ is the neutron star radius, $r$ is the orbital separation\cite{notation}), which is negligible except when $r$ is smaller than a few stellar radii. Since $R_o\simeq 5M$ for a typical neutron star of mass $1.4M_\odot$ and radius $10$ km, the tidal effect is essentially a (post)$^5$-Newtonian correction \cite{quad}. The expression for the phase error induced by the tidal effect is given in Ref.\cite{LRS94}. The fact that the evolution of the binary system as it sweeps through the low frequency band of the detector is insensitive to finite-size effects means that the measurement of the inspiral waveform will allow us to probe cleanly into the intricate structure of general relativity, and to test whether general relativity is the correct theory of gravity \cite{lucsathya,cliffscalar}. Moreover, some of the parameters of the binary system, such as the masses of the stars, can be determined with reasonable accuracy\cite{finnchernoff,cutlerflanagan}. However, the waveform's lack of dependence on the finite size of the objects during the most of the adiabatic inspiral also implies that information about the internal structure of the neutron star is only imprinted on the radiation emitted just prior to coalescence when the orbital radius is small. Indeed at small orbital separations, tidal effects are expected to be very important. In a purely {\it Newtonian} analysis, the interaction potential between star $M'$ and the tide-induced quadrupole of $M$, $V_{tide}\sim -M'^2R_o^5/r^6$, increases with decreasing $r$. The potential becomes so steep that a dynamical instability develops, accelerating the coalescence at small orbital radius. This Newtonian instability has been fully explored using semi-analytic models in Ref.\cite{LRS94} and Ref.\cite{LS95} (hereafter referred to as LS). It has also been examined numerically in Refs.\cite{Rasio,Centrella}. However, a purely Newtonian treatment of the binary at small separation is clearly not adequate, as general relativistic effects will also be important in this regime; and general relativistic effects can also make the orbit unstable. For example, a test particle in circular orbit around a Schwarzschild black hole will experience an ``innermost stable circular orbit'' at $r_{\rm isco}=6M$ (or $5M$ in harmonic coordinates). This unstable behavior is caused by higher-order relativistic corrections included in the Schwarzschild geodesic equations of motion. For computing the orbital evolution of two neutron stars of comparable mass near coalescence, the test-mass limit is obviously inadequate. In order to explore the orbital instability for such systems, Kidder, Will and Wiseman \cite{KWW93} (hereafter referred as KWW) developed {\it hybrid} equations of motion. These equations augment the the Schwarzschild geodesic equations of motion with the finite-mass terms of the (post)$^{5/2}$-Newtonian equations of motion. Including these finite mass terms in the equation of motion moved the innermost-stable-circular-orbit radius farther out (in units of the total mass).\footnote{See Wex and Sch\"afer\cite{Wex93} for a critique and an alternative construction. Their post-Newtonian calculation suggests that the innermost stable orbit may occur at an even greater separation.} In this paper, we augment the hybrid equations with contributions due to the tidal deformation of the stars. In a nutshell, the work presented here combines the Newtonian tidal analysis of LS \cite{LS95} with the relativistic point-mass analysis of KWW \cite{KWW93} to yield a more complete picture of the neutron-star coalescence prior to merging. We note, that unlike a test-particle around a Schwarzschild black hole, the very notion of ``innermost stable circular orbit'' is poorly defined for objects of comparable mass. After all, in the relativistic regime the binary orbit will be decaying rapidly due to radiation reaction; thus the orbit is not circular, but rather a decaying spiral. In order to give a semi-quantitative definition of ``innermost stable circular orbit'' we use the artifice of ``shutting off'' all the dissipative terms in the equation of motion and looking for the point where the solutions of the remaining non-dissipative equations become dynamically unstable. The use of hybrid equations of motion augmented with the tidal terms allows us to map out the dependence of the critical radius $r_{\rm isco}$, or the corresponding orbital frequency $f_{\rm isco}$, for a wide range of allowed neutron star equations of state (parametrized by radius and effective polytropic index; see Figure 1). We believe that clearing up such dependence is important, and this analysis provides a benchmark with which comparisons can be made with future numerical results. {\it Indeed, an important point we wish to make in this paper is that neither relativistic (post Newtonian) effects nor Newtonian tidal effects can be neglected near the instability limit, and the critical frequency can be much lower than the value obtained when only one of these effects are included.} \begin{figure}[t] \special{hscale=42 vscale=42 hoffset=0.0 voffset=-285.0 angle=0.0 psfile=fig1.eps} \vspace*{2.8in} \caption[Fig. 1]{ The critical orbital frequency (at the inner-most stable orbit) as a function of the ratio of the neutron star radius $R_o$ and mass $M$. The solid curves show the results including both relativistic and tidal effects (the lower curve is for $\Gamma=3$ while the upper one is for $\Gamma=2$), the dashed curves are the Newtonian limit given by Eqs.~(8)-(9). The vertical line corresponds to $R_o/M=9/4$, the minimum value for any physical neutron star. The insert is a close-up for the nominal range of $R_o/M=4-8$ as given by all the available nuclear EOS's. Two curves within the insert should bracket all the physical values of $f_{\rm isco}$.} \end{figure} The main results of our analysis are summarized in Figure 2 and 3. Figure 2 shows that the rate of radial infall for stars near coalescence is substantially underestimated if one models the coalescence as a Newtonian circular-orbit decaying solely under the influence of radiation reaction (top dotted curve). In other words, the rate of coordinate infall is substantially enhanced by the non-dissipative terms. Somewhat more relevant for observational purposes, Figure 3 shows that the number of orbits (or gravitational wave cycles) per logarithmic frequency interval is substantially reduced by the unstable collapse of the orbit. Both plots show modest sensitivity to the equation of state. \begin{figure}[t] \special{hscale=52 vscale=55 hoffset=-42.0 voffset=-365.0 angle=0.0 psfile=fig2.eps} \vspace*{3.1in} \caption[Fig. 2]{ The radial infall coordinate velocity during binary coalescence, with $M=M'=1.4M_\odot$, $R_o/M=5$, $\Gamma=3$, all calculated using $2.5PN$ radiation reaction. The solid line is the result including relativistic and tidal effects, the short-dashed line includes only tidal effects, the long-dashed line includes only relativistic effects. The dotted line is the point mass ``Newtonian'' result. } \end{figure} \begin{figure}[t] \special{hscale=52 vscale=55 hoffset=-50.0 voffset=-360.0 angle=0.0 psfile=fig3.eps} \vspace*{3.1in} \caption[Fig. 3]{ The number of orbits the binary spends per logarithmic frequency. The labels are the same as in Fig.~2. } \end{figure} The remainder of the paper is organized as follows: In section II we present our equations of motion. In section III we examine the orbital stability using the non-dissipative portion of the equations of motion, and thus identify the location of the ``innermost stable circular orbit''. In section IV we include the dissipative terms that were omitted in the analysis of section III, and evolve the full equations of motion. In section V we briefly discuss the relevance of our results to numerical hydrodynamic calculations and to gravitational wave signal analysis. \section{\bf EQUATIONS OF MOTION INCLUDING TIDAL AND GENERAL RELATIVISTIC EFFECTS} Consider a binary containing two neutron stars of mass $M$ and $M'$, each obeying a polytropic equation of state $P=K\rho^\Gamma$. We use the compressible ellipsoid model for binary stars developed in LS\cite{LS95}. Basically, we model the tidally deformed neutron star as an ellipsoid, with internal density profile similar to that of a spherical polytrope. The dynamics of such a neutron star (so called Riemann-S ellipsoid) is characterized by the three principal axes ($a_1,a_2,a_3$ for star $M$ and $a_1',a_2',a_3'$ for star $M'$), the angular velocity ($\Omega$ and $\Omega'$) of the ellipsoidal figure about a principal axis (perpendicular to the orbital plane) and the internal motion of the fluid with uniform vorticity. The non-zero internal fluid motion is necessary because the binary neutron stars are not expected to corotate with the orbit due to rapid orbital decay and small viscosity\cite{Bildsten92,Kochanek92}. Although the Newtonian tidal interaction between the neutron stars can be treated exactly in the linear regime using mode decomposition\cite{Lai94}, the ellipsoid model has the advantage that it can be extended to the nonlinear regime at small orbital radii, when the tidal deformation of the star becomes significant. The Newtonian dynamical equations for the binary neutron stars as derived in LS include the familiar Newtonian ($1/r^2$) force-law for point-masses orbiting one another; they further contain Newtonian terms involving finite size (tidal) effects. A post-Newtonian treatment of the tidal problem would give the relativistic corrections to these terms, namely the standard point-mass, post-Newtonian corrections to the equations of motion, as well as relativistic corrections to the quadrupole moment and corrections due to higher moments the bodies. (See Appendix F of \cite{WW96}.) To insure that our equations of motion at least agree with the known post-Newtonian, point-mass equations we augment these Newtonian equations of motion with the hybrid equations of KWW. However, we use only the Newtonian equations to describe the evolution of the neutron stars structure ($a_i$ and $a_i'$) and the fluid motion (the figure rotation rate and the internal vorticity) within the stars. These are given by Eqs.~(2.18)-(2.22) of LS. In other words, we neglect the relativistic corrections to the fluid motion, self-gravity and tidal interaction. These corrections are secondary effects and should not modify the {\it orbital dynamics} appreciably (e.g., the Newtonian tidal interaction between the two stars scales approximately as $M^2R_o^5/r^6$, and its relativistic correction is of order $M/r$ smaller). As noted before, the tidal interaction enters the Newtonian potential as a correction of $O[(R_o/r)^5]\sim O[(M_t/r)^5]$, in effect, as a (post)$^5$-Newtonian term. Therefore, by not including relativistic corrections to these tidal terms, we are merely omiting terms which are of (post)$^6$-Newtonian order. In fact, the largest error comes from neglecting the post-Newtonian correction (of order $M/R_o$) to the internal stellar structure (see Sec.~III.B for an estimate of its effect on $f_{\rm isco}$). The relativistic corrections to the orbital motion, however, are very important. Our equations of orbital motion can be assembled from Eqs.~(2.23)-(2.24) of LS and from Eqs.~(1.2)-(1.3) of KWW: \begin{eqnarray} \ddot r &=& r{\dot\theta}^2-{M_t\over r^2}(A_H+B_H\dot r)\nonumber\\ &&-{3\kappa_n\over 10}{M_t\over r^4}\left[a_1^2(3\cos^2\alpha-1) +a_2^2(3\sin^2\alpha-1)-a_3^2\right] \nonumber\\ &&-{3\kappa_n'\over 10}{M_t\over r^4}\left[a_1'^2(3\cos^2\alpha'-1) +a_2'^2(3\sin^2\alpha'-1)-a_3'^2\right] \nonumber\\ &&-{M_t\over r^2}(A_{5/2}+B_{5/2}\dot r) -{32\over 5}r\bigl[\Omega^5(I_{11}-I_{22})\sin 2\alpha \nonumber\\ &&+\Omega'^5(I_{11}'-I_{22}')\sin 2\alpha'\bigr], \\ \ddot\theta &=& -{2\dot r\dot\theta\over r} -{M_t\over r^2}B_H\dot\theta\nonumber\\ &&-{3\kappa_n\over 10}{M_t\over r^5}(a_1^2-a_2^2)\sin 2\alpha -{3\kappa_n'\over 10}{M_t\over r^5}(a_1'^2-a_2'^2)\sin 2\alpha' \nonumber\\ &&-{M_t\over r^2}B_{5/2}\dot\theta -{32\over 5}\bigl[\Omega^5(I_{11}-I_{22})\cos 2\alpha \nonumber\\ &&+\Omega'^5(I_{11}'-I_{22}')\cos 2\alpha'\bigr], \end{eqnarray} where $M_t=M+M'$ is the total mass, $\alpha$ ($\alpha'$) is the misalignment angle between the tidal bulge of $M$ ($M'$) and the line joing the two masses, $\kappa_n,\kappa_n'$ are dimensionless structure constants depending on the mass concentration within the stars. In Eqs.~(1) and (2) the last two lines contain the ``dissipative'' terms due to gravitational radiation reaction. The quantities $A_H$, $B_H$, $A_{5/2}$, $B_{5/2}$, which include the ``hybrid'' corrections to the equation of motion, are given by \begin{eqnarray} A_H &=& {1-M_t/r\over (1+M_t/r)^3}-\left[{2-M_t/r\over 1-(M_t/r)^2}\right]{M_t\over r}\dot r^2+v^2\nonumber\\ &&-\eta\left(2{M_t\over r}-3v^2+{3\over 2}\dot r^2\right) +\eta\biggl[{87\over 4}\left({M_t\over r}\right)^2\nonumber\\ &&+(3-4\eta)v^4+{15\over 8}(1-3\eta)\dot r^4 -{3\over 2}(3-4\eta)v^2\dot r^2\nonumber\\ &&-{1\over 2}(13-4\eta){M_t\over r}v^2 -(25+2\eta){M_t\over r}\dot r^2\biggr] \\ B_H &=& -\left[{4-2M_t/r\over 1-(M_t/r)^2}\right]\dot r +2\eta\dot r-{1\over 2}\eta\dot r\biggl[(15+4\eta)v^2\nonumber\\ &&-(41+8\eta){M_t\over r}-3(3+2\eta)\dot r^2\biggr], \\ A_{5/2} &=& -{8\over 5}\eta{M_t\over r}\dot r\left(18v^2 +{2\over 3}{M_t\over r}-25\dot r^2\right), \\ B_{5/2} &=& {8\over 5}\eta{M_t\over r}\left(6v^2 -2{M_t\over r}-15\dot r^2\right) \; , \end{eqnarray} where $v^2=\dot r^2+r^2\dot\theta^2$, $\eta=\mu/M_t$ and $\mu=MM'/M_t$. Also the multipoles moments can be expressed as \begin{equation} I_{ii}=\kappa_nMa_i^2/5,~~~~~ I_{ii}'=\kappa_n'M'a_i'^2/5 \; . \end{equation} Note that in Eqs.~(1)-(2), we have also included the leading order radiation reaction forces due to tidal deformation. Admittedly, this is not a consistent post-Newtonian expansion of the true equations of motion; however it is correct in several important limiting cases: (i) In the limit that $a_i \rightarrow 0$ and $a_i' \rightarrow 0$ and the limit $\eta \rightarrow 0$, we recover the {\it exact} Schwarzschild equation of motion. (ii) In the point-mass limit ($a_i\rightarrow 0$ and $a_i'\rightarrow 0$) we recover the hybrid equations given in KWW. KWW presented an argument that suggested that the higher-order, $\eta$-dependent, (post)$^3$-Newtonian -- as yet uncalculated -- corrections to these equations have only a modest effect on the equations of motion. See Figure 6 of Ref.\cite{KWW93}. However, until these terms are calculated it is unclear just how large an effect they will have on the location of the innermost stable orbit. (iii) In the non-relativistic limit we recover the equations of motion given in LS. These equations contain the dominant contributions to the equations of motion due to the finite sizes of the objects. Note that although Eqs.~(1) and (2) make reference to the orbital radius $r$, we are always aware that this is a gauge dependent quantity and of little meaning for a distant observer. Observationally, the more meaningful quantity is the orbital frequency as measured by distant observers, and we shall use frequency rather the radius in presenting most of our results. \section{\bf INSTABILITY OF THE NON-DISIPATIVE EQUATIONS OF MOTION} \subsection{Method to Determine the Stability Limit} We now form a set of non-dissipative equations of motion, by simply discarding the gravitational radiation reaction terms given in the last two lines of Eq.~(1) and Eq.~(2). These non-dissipative dynamical equations admit equilibrium solutions, which are obtained by setting $\dot r=\ddot r=\ddot\theta=\dot\Omega_{orb}=\alpha=\alpha'=0$ as well as $\dot a_i=\dot a_i'=0$. For a given $r$, the evolution equations for the neutron star structure reduce to a set of algebraic equations for $a_i$ and $a_i'$, while the orbital equation (2) gives the orbital frequency $\Omega_{orb}$. These equations are solved using a Newton-Raphson method, yielding an equilibrium binary model. Thus a sequence of binary models parametrized by $r$ can be constructed. To determine the stability of the orbit of a binary model, we simply use the equilibrium parameters as initial conditions for our non-dissipative equations of motion. We add a small perturbation to the equilibrium model and let the system evolve. In this way we locate the critical point of the dynamical equations, corresponding to the dynamical stability limit of the equilibrium binary or the inner-most stable circular orbit: for $r>r_{\rm isco}$, the binary is stable, and the system oscillates with small amplitude about the initial configuration; for $r<r_{\rm isco}$, the binary is unstable, and the perturbation grows, leading to the swift merger of the neutron stars even in the absence of dissipation. \subsection{Results} For concreteness, we present results only for binary neutron stars with equal masses ($M=M'$), both having zero spin at large orbital separation, although our equations are adequate to treat the most general cases\cite{LS95}. The polytropic relation $P=K\rho^\Gamma$ provides a useful parametrization to the most general realistic nuclear equation of state (EOS). Since the radius $R_o$ of the nonrotating neutron star of mass $M$ is uniquely determined by $K$ and $\Gamma$, we can alternatively use $R_o/M$ and $\Gamma$ to characterize the EOS. For a canonical neutron star with mass $M=1.4M_\odot$, all EOS tabulated in\cite{Arnett77} give $R_o/M$ in the range of $4-8$, while modern microscopic nuclear calculations typically give $R_o/M=5$\cite{Wiringa88}. For a given $R_o/M$, the polytropic index $\Gamma$ specifies the mass concentration within the star. Except for extreme neutron star masses ($M\lo 0.5M_\odot$ or $M\go 1.8M_\odot$) typical values of $\Gamma$ lie in the range of $\Gamma=2-3$ \cite{LRS94}. In Table I, we list the physical properties of the equilibrium binary neutron stars at the dynamical stability limit for several values of $R_o/M$ and $\Gamma=3$. In Figure 1, the orbital frequency $f_{\rm isco}$ is shown as a function of $R_o/M$ for $\Gamma=2$ and $\Gamma=3$. Clearly, in the limit of $R_o/M\rightarrow 0$, $f_{\rm isco}$ approachs the point mass result $f_{\rm isco}=697M_{1.4}^{-1}$ Hz obtained in KWW\footnote{KWW\cite{KWW93} give a correct expression for $r_{\rm isco}/M_t$, but incorrectly give $f_{\rm isco}$=710 Hz due to a numerical error.}. In the non-relativistic limit we recover the pure Newtonian result\cite{LRS94,LS95}: \begin{eqnarray} f_{\rm isco} &=& 657M_{1.4}^{-1}(5M/R_o)^{3/2}~({\rm Hz}) ~~~~~(\Gamma=3),\\ f_{\rm isco} &=& 766M_{1.4}^{-1}(5M/R_o)^{3/2}~({\rm Hz}) ~~~~~(\Gamma=2). \end{eqnarray} {\it For typical neutron star radius $R_o/M=5$, the critical frequency ranges from $488$ Hz (for $\Gamma=3$) to $540$ Hz (for $\Gamma=2$), while both the pure Newtonian (with tides) calculation and the pure point-mass hybrid calculation give a result $30-40\%$ larger}. There are two physical causes for the reduction in $f_{\rm isco}$: (i) The binary becomes unstable at larger orbital separation due to the steepening of the interaction potential from both tidal and relativistic effects; (ii) For a given orbital radius (itself a gauge dependent quantity), the post-Newtonian orbital frequency as measured by an observer at infinity is smaller than the Newtonian orbital frequency\footnote{In the case of equal masses, at first post-Newtonian order $\Omega_{orb} = \Omega_{Kepler}[1 - (11/8)(M_t/r)]$. See Ref.\cite{bdiww}.}. We conclude that to neglect either the tidal effects or the relativistic effects can lead to large error in the estimated critical frequency. Except for the intrinsic uncertainties associated with the hybrid equations of motion\cite{KWW93,Wex93}, the main uncertainty in our determination of $f_{\rm isco}$ comes from neglecting post-Newtonian corrections to (i) the stellar structure and (ii) the the tidal potential. The first correction {\it decreases} the tide-induced quadrupole; the fractional change is of order $-M/R_o$. The second {\it increases} the quadrupole by a fraction of order $M'/r$. We can estimate how much $f_{\rm isco}$ is modified by these two corrections. In Newtonian theory, $r_{\rm isco}$ is approximately determined by the condition $MM'/r\sim M'^2R_o^5/r^6$. Including the relativistic corrections this condition becomes $MM'/r\sim M'^2R_o^5/r^6(1-\delta)$, where $\delta\sim [O(M/R_o)-O(M'/r)]\lo 20\%$. Thus the change in $f_{\rm isco}$ due to these two effects is $\Delta f_{\rm isco}/f_{\rm isco}\simeq 0.3\,\delta\lo 6\%$, i.e., the critical frequency increases by a few percent\cite{Wilsonnote}. As emphasized in Sec.~I, the critical radius (or critical frequency), at which the non-dissipative equations become dynamically unstable, is meaningful only in the sense that when $r<r_{\rm isco}$, the binary will coalesce on dynamical (orbital) timescale even in the absence of dissipation. In the realistic situation, the dissipative radiation reaction forces will also be rapidly driving the binary to coalescence. Therefore to determine the significance of the dynamical instability we must compute the orbital evolution with the full equations of motion -- including the radiation reaction. \section{\bf ORBITAL EVOLUTION PRIOR TO MERGER} We now include the dissipative radiation reaction forces in our analysis. In this case the plunge will be driven by both the dissipative, as well as the non-dissipative effects associated with the steepening potential (both the tidal potential and the relativistic potential). But what effect is dominant? In order to numerically investigate this question, we choose a specific system with $M=M'=1.4M_\odot$, $R_o/M=5$ and $\Gamma=3$. The orbital evolution begins when the stars are well outside the innermost stable circular orbit limit. We consider four different inspiral scenarios. (i) A purely dissipative inspiral: a system of point masses subject only to a Newtonian ($1/r^2$) force and (post)$^{5/2}$-Newtonian radiation reaction force. In this case the infall rate is given by $v_r=\dot r=-(64/5)\eta (M_t/r)^3$. [Specifically, we set $A_H = 1$ and $B_H = a_i = a_i' = I_{kk} = I_{kk}' = 0$ in Eqs.~(1)-(2).] This is depicted by the dotted curve in Figure 2 and 3. (ii) A purely relativistic plunge in which we neglect the tidal effects [Specifically we set $a_i=a_i'= I_{kk} = I_{kk}' =0$ in Eqs.~(1)-(2)]. This relativistic case is depicted by the long-dashed curve in Figure 2 and 3. (iii) A tidally enhanced plunge: we include only the Newtonian terms in Eqs.~(1)-(2) and the radiation reaction force. [Specifically we set $A_H = 1$ and $B_H = 0$ in Eqs.~(1)-(2).] This case is depicted by the short-dashed curve in Figure 2 and 3. (iv) Finally, we evolve the complete dynamical equations including all terms in Eqs.~(1)-(2); this is depicted by the solid curve in Figure 2 and 3. Each intergration is terminated when the surfaces of the stars touch, i.e., at $r\simeq 2a_1$ (for the point-mass problem, the calculation is terminated at $r\simeq 2R_o$). In Figure 2 we clearly see that the non-dissipative effects -- tidal and relativistic -- substantially increase the rate of infall. The radial velocity at binary contact is comparable to the tangential velocity. We also note that the radial coordinate velocity is a gauge dependent quantity; therefore our only intent in using it in Figure 2 is to convey the general trend that the rate of infall is enhanced by the dynamical instability. Figure 3 shows the number of orbits the binary spends per logarithmic frequency. In the simplest point-mass, Newtonian-plus-radiation-reaction case [case (i) above], the result can be calculated analytically \begin{eqnarray} dN_{orb}/d\ln f_{orb} &=&(5/192\pi)\mu^{-1}M_t^{-2/3} (2\pi f_{orb})^{-5/3}\nonumber\\ &=&1.95\times 10^5(M_{1.4}f_{orb}/{\rm Hz})^{-5/3}, \end{eqnarray} which gives $6$ cycles at $f_{gw}\simeq 2f_{orb}=1000$ Hz. In contrast, the tidal and relativistic effects reduce this number to less than $2$. Figure 4 shows the wave energy emitted around a given frequency, $dE_{gw}/d\ln f_{orb}=(\Omega_{orb}/\dot\Omega_{orb})\dot E_{gw}$, where $\dot E_{gw}$ is calculated using the simple quadrupole radiation formula. The Newtonian plus radiation reaction result is $dE_{gw}/d\ln f_{orb}=1.63\times 10^{-3}(f_{orb}/ {\rm Hz})^{2/3}M^2/R_o$. We see that the radiation power near contact becomes much smaller. Note that $dE_{gw}/d\ln f_{orb}$ calcuated in this way is not exactly the energy power spectrum, which must be obtained from the Fourier transform of the waveform\cite{Kennefick96}; however, it does provide a semi-quantitative feature of the full analysis; in particular, the dip in the $dE_{gw}/d\ln f_{orb}$ curve around $600$ Hz results from the dynamical instability of the orbit (see also Refs.\cite{Centrella,Ruffert}, although the calculations presented there are purely Newtonian). \begin{figure}[t] \special{hscale=52 vscale=55 hoffset=-50.0 voffset=-375.0 angle=0.0 psfile=fig4.eps} \vspace*{3.1in} \caption[Fig. 4]{ The quadrupolar gravitational energy emitted near a given frequency. The labels are the same as in Fig.~2. } \end{figure} \section{\bf DISCUSSION} A number of authors have tried to define and locate the innermost stable circular orbit for relativistic coalescing systems of comparable masses \cite{Wex93,Wilson95,clarkeardley,blackburndetweiler,cook} in order to characterize the final moments of a binary coalescence (See \cite{eardleyhirschnamm} for a discussion). The results of the various analyses are not converging to an agreed answer. Obviously, the precise nature of the final coalescence of two neutron stars will only be determined by a numerical simulation using full general-relativistic hydrodynamics. However, the present analysis does point to two intersting features to look for in a full numerical treatment: (i) To get even a qualitative picture of the coalescence, it is necessary to begin the numerical evolution when the stars are still well separated, {\it i.e.} before the onset of the orbital dynamical instability. The instability -- the plunge -- causes the coalescence to proceed much more swiftly than a coalescence driven solely by radiation reaction; thus the actual coalescence may differ qualitatively from one computed with a simple radiation-reaction driven inspiral. The final coalescence may be more of a splat, than the slow winding together of the stars\cite{Nakamura}. (ii) The instability results from both the tidal effects and the relativistic corrections in the equations of motion. KWW showed that there is no instability in the {\it first} post-Newtonian relativistic equations of motion; the instability does not show up until at least second post-Newtonian order. Therefore, in order for numerical simulations to see the effects of the relativistic unstable plunge, it will probably require the use of at least second-order, post-Newtonian hydrodynamic code. KWW also showed that the location of the dynamic instability does not converge very rapidly as one increases the post-Newtonian order of the approximation. [This fact led KWW to the introduce the hybrid equations of motion.] Thus, to get even a qualitatively accurate evolution of the binary near coalescence, it may be necessary to use a full general relativistic hydrodynamic treatment of the coalescence problem, and begin the evolution when the stars are still well separated. As we have shown, the dynamical instability in the equation of motion will, in effect, cut off the chirping waveform. The frequency of the cut-off is somewhat dependent on the neutron star equation of state. Only in this late stage of the evolution does the equation of state leave a tell-tale sign in the emitted waveform. However, devising a strategy to dig this information from the detector output requires further consideration. Most detection/measurement strategies for coalescing binaries involve integrating template waveforms against long stretches (8000 orbits!) of raw output data, the idea being that one can detect/measure a relatively low amplitude signal by integrating for a long time. Looking for the signature of this very late stage plunge is precisely the opposite: we are looking at the waveform just before coalescence when the amplitude is fairly strong, but the plunge is of fairly short in duration. So answering questions about the plunge (such as, at what orbital frequency did it begin?) requires measuring a relatively large amplitude, but short-duration, effect. Clearly, analysis of such events will require a different detection strategy\cite{Kennefick96}. \acknowledgments We thank Kip Thorne for useful discussions. This work has been supported by NSF Grants AST-9417371, PHY-9424337 and NASA Grant NAGW-2756 to Caltech. DL also acknowledges support of the Richard C. Tolman Fellowship at Caltech.
proofpile-arXiv_065-481
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} \bigskip Over the past decade, beginning with the measurement of nucleon spin--polarized structure function, $g_{1}(x,Q^2)$ by the EMC \cite{emc88} at CERN and most recently with the spin--structure function $g_2(x,Q^2)$ in the E143 experiment \cite{slac96} at SLAC, a wealth of information has been gathered on the spin--polarized structure functions of the nucleon and their corresponding sum rules (see in addition \cite{smc93}, \cite{smc94a}, \cite{smc94b}, \cite{slac93}, \cite{slac95a}, \cite{slac95b}). Initially the analysis of these experiments cast doubt on the non--relativistic quark model \cite{kok79} interpretations regarding the spin content of the proton. By now it is firmly established that the quark helicity of the nucleon is much smaller than the predictions of that model, however, many questions remain to be addressed concerning the spin structure. As a result there have been numerous investigations within models for the nucleon in an effort to determine the manner in which the nucleon spin is distributed among its constituents. One option is to study the axial current matrix elements of the nucleon such as $\langle N |{\cal A}_{\mu}^{i}|N\rangle = 2\Delta q_{i}S_{\mu}$, which, for example, provide information on the nucleon axial singlet charge \begin{eqnarray} g_{A}^{0}&=&\langle N |{\cal A}_{3}^{0}|N\rangle\ = \left( \Delta u + \Delta d + \Delta s\right)= \Gamma_{1}^{p} (Q^2) + \Gamma_{1}^{n} (Q^2) \ . \end{eqnarray} Here $\Delta q$ are the axial charges of the quark constituents and $\Gamma_{1}^{N} (Q^2)=\int_{0}^{1} dx g_{1}^{N}(x,Q^2)$ is the first moment of the longitudinal nucleon spin structure function, $g_1^N\left(x,Q^{2}\right)$. Of course, it is more illuminating to directly compute the longitudinal and transverse nucleon spin--structure functions, $g_{1}\left(x,Q^2\right)$ and $g_{T}(x,Q^2)=g_{1}(x,Q^2)+g_{2}(x,Q^2)$, respectively as functions of the Bjorken variable $x$. We will calculate these structure functions within the Nambu--Jona--Lasinio (NJL) \cite{Na61} chiral soliton model \cite{Re88}. Chiral soliton models are unique both in being the first effective models of hadronic physics to shed light on the so called ``proton--spin crisis" by predicting a singlet combination in accord with the data \cite{br88}, and in predicting a non--trivial strange quark content to the axial vector current of the nucleon \cite{br88}, \cite{pa89}, \cite{jon90}, \cite{blo93}; about $10-30\%$ of the down quarks (see \cite{wei96a} and \cite{ell96} for reviews). However, while the leading moments of these structure functions have been calculated within chiral soliton models, from the Skyrme model \cite{Sk61}, \cite{Ad83} and its various vector--meson extensions, to models containing explicit quark degrees of freedom such as the (NJL) model \cite{Na61}, the nucleon spin--structure functions themselves have not been investigated in these models. Soliton model calculations of structure functions were, however, performed in Friedberg-Lee \cite{frie77} and color-dielectric \cite{nil82} models. In addition, structure functions have extensively been studied within the framework of effective quark models such as the bag--model \cite{Ch74}, and the Center of Mass bag model \cite{So94}. These models are confining by construction but they neither contain non--perturbative pseudoscalar fields nor are they chirally symmetric\footnote{In the cloudy bag model the contribution of the pions to structure functions has at most been treated perturbatively \cite{Sa88}, \cite{Sc92}.}. To this date it is fair to say that many of the successes of low--energy effective models rely on the incorporation of chiral symmetry and its spontaneous symmetry breaking (see for e.g. \cite{Al96}). In this article we therefore present our calculation of the polarized spin structure functions in the NJL chiral soliton model \cite{Re89}, \cite{Al96}. Since in particular the static axial properties of the nucleon are dominated by the valence quark contribution in this model it is legitimate to focus on the valence quarks in this model. At the outset it is important to note that a major difference between the chiral soliton models and models previously employed to calculate structure functions is the form of the nucleon wave--function. In the latter the nucleon wave--function is a product of Dirac spinors while in the former the nucleon appears as a collectively excited (topologically) non--trivial meson configuration. As in the original bag model study \cite{Ja75} of structure functions for localized field configurations, the structure functions are most easily accessible when the current operator is at most quadratic in the fundamental fields and the propagation of the interpolating field can be regarded as free. Although the latter approximation is well justified in the Bjorken limit the former condition is difficult to satisfy in soliton models where mesons are fundamental fields ({\it e.g.} the Skyrme model \cite{Sk61}, \cite{Ad83}, the chiral quark model of ref. \cite{Bi85} or the chiral bag model \cite{Br79}). Such model Lagrangians typically possess all orders of the fundamental pion field. In that case the current operator is not confined to quadratic order and the calculation of the hadronic tensor (see eq. (\ref{deften}) below) requires drastic approximations. In this respect the chirally invariant NJL model is preferred because it is entirely defined in terms of quark degrees of freedom and formally the current possesses the structure as in a non--interacting model. This makes the evaluation of the hadronic tensor feasible. Nevertheless after bosonization the hadronic currents are uniquely defined functionals of the solitonic meson fields. The paper is organized as follows: In section 2 we give a brief discussion of the standard operator product expansion (OPE) analysis to establish the connection between the effective models for the baryons at low energies and the quark--parton model description. In section 3 we briefly review the NJL chiral soliton. In section 4 we extract the polarized structure functions from the hadronic tensor, eq. (\ref{had}) exploiting the ``valence quark approximation". Section 5 displays the results of the spin--polarized structure functions calculated in the NJL chiral soliton model within this approximation and compare this result with a recent low--renormalization point parametrization \cite{Gl95}. In section 6 we use Jaffe's prescription \cite{Ja80} to impose proper support for the structure function within the interval $x\in \left[0,1\right]$. Subsequently the structure functions are evolved \cite{Al73}, \cite{Al94}, \cite{Ali91} from the scale characterizing the NJL--model to the scale associated with the experimental data. Section 7 serves to summarize these studies and to propose further explorations. In appendix A we list explicit analytic expressions for the isoscalar and isovector polarized structure functions. Appendix B summarizes details on the evolution of the twist--3 structure function, ${\overline{g}}_2\left(x,Q^2\right)$. \bigskip \section{DIS and the Chiral Soliton} \bigskip It has been a long standing effort to establish the connection between the chiral soliton picture of the baryon, which essentially views baryons as mesonic lumps and the quark parton model which regards baryons as composites of almost non--interacting, point--like quarks. While the former has been quite successful in describing static properties of the nucleon, the latter, being firmly established within the context of deep inelastic scattering (DIS), has been employed extensively to calculate the short distance or perturbative processes within QCD. In fact this connection can be made through the OPE. The discussion begins with the hadronic tensor for electron--nucleon scattering, \begin{eqnarray} W_{\mu\nu}(q)=\frac{1}{4\pi}\int d^4 \xi \ {\rm e}^{iq\cdot\xi} \langle N |\left[J_\mu(\xi),J^{\dag}_\nu(0)\right]|N\rangle\ , \label{deften} \end{eqnarray} where $J_\mu={\bar q}(\xi)\gamma_\mu {\cal Q} q(\xi)$ is the electromagnetic current, ${\cal Q}=\left(\frac{2}{3},\frac{-1}{3}\right)$ is the (two flavor) quark charge matrix and $|N\rangle$ refers to the nucleon state. In the DIS regime the OPE enables one to express the product of these currents in terms of the forward Compton scattering amplitude $T_{\mu\nu}(q)$ of a virtual photon from a nucleon \begin{eqnarray} T_{\mu\nu}(q)=i\int d^4 \xi \ {\rm e}^{iq\cdot\xi} \langle N |T\left(J_\mu(\xi)J^{\dag}_\nu(0)\right)|N\rangle\ , \label{im} \end{eqnarray} by an expansion on the light cone $\left(\xi^2 \rightarrow 0\right)$ using a set of renormalized local operators \cite{muta87}, \cite{rob90}. In the Bjorken limit the influence of these operators is determined by the twist, $\tau$ or the light cone singularity of their coefficient functions. Effectively this becomes a power series in the inverse of the Bjorken variable $x=-q^{2}/2P\cdot q$, with $P_\mu$ being the nucleon momentum: \begin{eqnarray} T_{\mu\nu}(q)\ =\sum_{n,i,\tau} \left(\frac{1}{x}\right)^{n}\ e_{\mu\nu}^{i}\left(q,P,S\right)\ C^{n}_{\tau,i}(Q^2/\mu^2,\alpha_s(\mu^2)){\cal O}^{n}_{\tau,i}(\mu^2) (\frac{1}{Q^2})^{\frac{\tau}{2}\ - 1}\ . \label{series} \end{eqnarray} Here the index $i$ runs over all scalar matrix elements, ${\cal O}_{\tau,i}^{n}(\mu^2)$, with the same Lorentz structure (characterized by the tensor, $e_{\mu\nu}^{i}$). Furthermore, $S^{\mu}$ is the spin of the nucleon, $\left(S^2=-1\ , S\cdot P\ =0\right)$ and $Q^2=-q^2 > 0$. As is evident, higher twist contributions are suppressed by powers of $1/{Q^2}$. The coefficient functions, $C^{n}_{\tau,i}(Q^2/\mu^2,\alpha_s(\mu^2))$ are target independent and in principle include all QCD radiative corrections. Their $Q^2$ variation is determined from the solution of the renormalization group equations and logarithmically diminishes at large $Q^2$. On the other hand the reduced--matrix elements, ${\cal O}_{\tau,i}^{n}(\mu^2)$, depend only on the renormalization scale $\mu^2$ and reflect the non--perturbative properties of the nucleon \cite{ans95}. The optical theorem states that the hadronic tensor is given in terms of the imaginary part of the virtual Compton scattering amplitude, $W_{\mu\nu}=\frac{1}{2\pi}{\rm Im}\ T_{\mu\nu}$. From the analytic properties of $T_{\mu\nu}(q)$, together with eq. (\ref{series}) an infinite set of sum rules result for the form factors, ${\cal W}_{i}\left(x,Q^2\right)$, which are defined via the Lorentz covariant decomposition $W_{\mu\nu}(q)=e_{\mu\nu}^{i}{\cal W}_i\left(x,Q^2\right)$. These sum rules read \begin{eqnarray} \int^{1}_{0}dx\ x^{n-1}\ {\cal W}_{i}\left(y,Q^2\right)&=& \sum _{\tau}\ C^{n}_{\tau,i}\left(Q^2/\mu^2,\alpha_s(\mu^2)\right) {\cal O}_{\tau,i}^{n}(\mu^2) (\frac{1}{Q^2})^{\frac{\tau}{2}\ - 1}\ . \label{ope} \end{eqnarray} In the {\em impulse approximation} (i.e. neglecting radiative corrections) \cite{Ja90,Ji90,Ja91} one can directly sum the OPE gaining direct access to the structure functions in terms of the reduced matrix elements ${\cal O}_{\tau,i}^{n}(\mu^2)$. When calculating the renormalization--scale dependent matrix elements, ${\cal O}_{\tau,i}^{n}(\mu^2)$ within QCD, $\mu^2$ is an arbitrary parameter adjusted to ensure rapid convergence of the perturbation series. However, given the difficulties of obtaining a satisfactory description of the nucleon as a bound--state in the $Q^2$ regime of DIS processes it is customary to calculate these matrix elements in models at a low scale $\mu^2$ and subsequently evolve these results to the relevant DIS momentum region of the data employing, for example, the Altarelli--Parisi evolution \cite{Al73}, \cite{Al94}. In this context, the scale, $\mu^2 \sim \Lambda_{QCD}^{2}$, characterizes the non--perturbative regime where it is possible to formulate a nucleon wave--function from which structure functions are computed. Here we will utilize the NJL chiral--soliton model to calculate the spin--polarized nucleon structure functions at the scale, $\mu^2$, subsequently evolving the structure functions according to the Altarelli--Parisi scheme. This establishes the connection between chiral soliton and the parton models. In addition we compare the structure functions calculated in the NJL model to a parameterization of spin structure function \cite{Gl95} at a scale commensurate with our model. \bigskip \section{The Nucleon State in the NJL Model} \bigskip The Lagrangian of the NJL model reads \begin{eqnarray} {\cal L} = \bar q (i\partial \hskip -0.5em / - m^0 ) q + 2G_{\rm NJL} \sum _{i=0}^{3} \left( (\bar q \frac {\tau^i}{2} q )^2 +(\bar q \frac {\tau^i}{2} i\gamma _5 q )^2 \right) . \label{NJL} \end{eqnarray} Here $q$, $\hat m^0$ and $G_{\rm NJL}$ denote the quark field, the current quark mass and a dimensionful coupling constant, respectively. When integration out the gluon fields from QCD a current--current interaction remains, which is meditated by the gluon propagator. Replacing this gluon propagator by a local contact interaction and performing the appropriate Fierz--transformations yields the Lagrangian (\ref{NJL}) in leading order of $1/N_c$ \cite{Re90}, where $N_c$ refers to the number of color degrees of freedom. It is hence apparent that the interaction term in eq. (\ref{NJL}) is a remnant of the gluon fields. Hence gluonic effects are included in the model described by the Lagrangian (\ref{NJL}). Application of functional bosonization techniques \cite{Eb86} to the Lagrangian (\ref{NJL}) yields the mesonic action \begin{eqnarray} {\cal A}&=&{\rm Tr}_\Lambda\log(iD)+\frac{1}{4G_{\rm NJL}} \int d^4x\ {\rm tr} \left(m^0\left(M+M^{\dag}\right)-MM^{\dag}\right)\ , \label{bosact} \\ D&=&i\partial \hskip -0.5em /-\left(M+M^{\dag}\right) -\gamma_5\left(M-M^{\dag}\right)\ . \label{dirac} \end{eqnarray} The composite scalar ($S$) and pseudoscalar ($P$) meson fields are contained in $M=S+iP$ and appear as quark--antiquark bound states. The NJL model embodies the approximate chiral symmetry of QCD and has to be understood as an effective (non--renormalizable) theory of the low--energy quark flavor dynamics. For regularization, which is indicated by the cut--off $\Lambda$, we will adopt the proper--time scheme \cite{Sch51}. The free parameters of the model are the current quark mass $m^0$, the coupling constant $G_{\rm NJL}$ and the cut--off $\Lambda$. Upon expanding ${\cal A}$ to quadratic order in $M$ these parameters are related to the pion mass, $m_\pi=135{\rm MeV}$ and pion decay constant, $f_\pi=93{\rm MeV}$. This leaves one undetermined parameter which we choose to be the vacuum expectation value $m=\langle M\rangle$. For apparent reasons $m$ is called the constituent quark mass. It is related to $m^0$, $G_{\rm NJL}$ and $\Lambda$ via the gap--equation, {\it i.e.} the equation of motion for the scalar field $S$\cite{Eb86}. The occurrence of this vacuum expectation value reflects the spontaneous breaking of chiral symmetry and causes the pseudoscalar fields to emerge as (would--be) Goldstone bosons. As the NJL model soliton has exhaustively been discussed in recent review articles \cite{Al96}, \cite{Gok96} we only present those features, which are relevant for the computation of the structure functions in the valence quark approximation. The chiral soliton is given by the hedgehog configuration of the meson fields \begin{eqnarray} M_{\rm H}(\mbox{\boldmath $x$})=m\ {\rm exp} \left(i\mbox{\boldmath $\tau$}\cdot{\hat{\mbox{\boldmath $x$}}} \Theta(r)\right)\ . \label{hedgehog} \end{eqnarray} In order to compute the functional trace in eq. (\ref{bosact}) for this static configuration we express the Dirac operator (\ref{dirac}) as, $D=i\gamma_0(\partial_t-h)$ where \begin{eqnarray} h=\mbox{\boldmath $\alpha$}\cdot\mbox{\boldmath $p$}+m\ {\rm exp}\left(i\gamma_5\mbox{\boldmath $\tau$} \cdot{\hat{\mbox{\boldmath $x$}}}\Theta(r)\right)\ \label{hamil} \end{eqnarray} is the corresponding Dirac Hamiltonian. We denote the eigenvalues and eigenfunctions of $h$ by $\epsilon_\mu$ and $\Psi_\mu$, respectively. Explicit expressions for these wave--functions are displayed in appendix A. In the proper time regularization scheme the energy functional of the NJL model is found to be \cite{Re89,Al96}, \begin{eqnarray} E[\Theta]= \frac{N_C}{2}\epsilon_{\rm v} \left(1+{\rm sgn}(\epsilon_{\rm v})\right) &+&\frac{N_C}{2}\int^\infty_{1/\Lambda^2} \frac{ds}{\sqrt{4\pi s^3}}\sum_\nu{\rm exp} \left(-s\epsilon_\nu^2\right) \nonumber \\* && \hspace{1.5cm} +\ m_\pi^2 f_\pi^2\int d^3r \left(1-{\rm cos}\Theta(r)\right) , \label{efunct} \end{eqnarray} with $N_C=3$ being the number of color degrees of freedom. The subscript ``${\rm v}$" denotes the valence quark level. This state is the distinct level bound in the soliton background, {\it i.e.} $-m<\epsilon_{\rm v}<m$. The chiral angle, $\Theta(r)$, is obtained by self--consistently extremizing $E[\Theta]$ \cite{Re88}. States possessing good spin and isospin quantum numbers are generated by rotating the hedgehog field \cite{Ad83} \begin{eqnarray} M(\mbox{\boldmath $x$},t)= A(t)M_{\rm H}(\mbox{\boldmath $x$})A^{\dag}(t)\ , \label{collrot} \end{eqnarray} which introduces the collective coordinates $A(t)\in SU(2)$. The action functional is expanded \cite{Re89} in the angular velocities \begin{eqnarray} 2A^{\dag}(t)\dot A(t)= i\mbox{\boldmath $\tau$}\cdot\mbox{\boldmath $\Omega$} \ . \label{angvel} \end{eqnarray} In particular the valence quark wave--function receives a first order perturbation \begin{eqnarray} \Psi_{\rm v}(\mbox{\boldmath $x$},t)= {\rm e}^{-i\epsilon_{\rm v}t}A(t) \left\{\Psi_{\rm v}(\mbox{\boldmath $x$}) +\frac{1}{2}\sum_{\mu\ne{\rm v}} \Psi_\mu(\mbox{\boldmath $x$}) \frac{\langle \mu |\mbox{\boldmath $\tau$}\cdot \mbox{\boldmath $\Omega$}|{\rm v}\rangle} {\epsilon_{\rm v}-\epsilon_\mu}\right\}=: {\rm e}^{-i\epsilon_{\rm v}t}A(t) \psi_{\rm v}(\mbox{\boldmath $x$}). \label{valrot} \end{eqnarray} Here $\psi_{\rm v}(\mbox{\boldmath $x$})$ refers to the spatial part of the body--fixed valence quark wave--function with the rotational corrections included. Nucleon states $|N\rangle$ are obtained by canonical quantization of the collective coordinates, $A(t)$. By construction these states live in the Hilbert space of a rigid rotator. The eigenfunctions are Wigner $D$--functions \begin{eqnarray} \langle A|N\rangle=\frac{1}{2\pi} D^{1/2}_{I_3,-J_3}(A)\ , \label{nwfct} \end{eqnarray} with $I_3$ and $J_3$ being respectively the isospin and spin projection quantum numbers of the nucleon. \bigskip \section{Polarized Structure Functions in the NJL model} \bigskip The starting point for computing nucleon structure functions is the hadronic tensor, eq. (\ref{deften}). The polarized structure functions are extracted from its antisymmetric piece, $W^{(A)}_{\mu\nu}=(W_{\mu\nu}-W_{\nu\mu})/2i$. Lorentz invariance implies that the antisymmetric portion, characterizing polarized lepton--nucleon scattering, can be decomposed into the polarized structure functions, $g_1(x,Q^2)$ and $g_2(x,Q^2)$, \begin{eqnarray} W^{(A)}_{\mu\nu}(q)= i\epsilon_{\mu\nu\lambda\sigma}\frac{q^{\lambda}M_N}{P\cdot q} \left\{g_1(x,Q^2)S^{\sigma}+ \left(S^{\sigma}-\frac{q\cdot S}{q\cdot p}P^{\sigma}\right) g_2(x,Q^2)\right\}\ , \label{had} \end{eqnarray} again, $P_\mu$ refers to the nucleon momentum and $Q^2=-q^2$. The tensors multiplying the structure functions in eq. (\ref{had}) should be identified with the Lorentz tensors $e_{\mu\nu}^{i}$ in (\ref{series}). Contracting $W^{(A)}_{\mu\nu}$ with the longitudinal $\Lambda^{\mu\nu}_{L}$ and transverse $\Lambda^{\mu\nu}_{T}$ projection operators \cite{ans95}, \begin{eqnarray} \Lambda^{\mu\nu}_{L}&=&\frac{2}{b}\left\{2P\cdot qxS_{\lambda}+ \frac{1}{q\cdot S}\left[(q\cdot S)^{2}- \left(\frac{P\cdot q}{M}\right)^2\right] q_{\lambda}\right\}\ P_\tau \ \epsilon^{\mu\nu\lambda\tau }, \label{proj1}\\ \Lambda^{\mu\nu}_{T}&=&\frac{2}{b} \left\{\left[\left(\frac{P\cdot q}{M}\right)^2+2P\cdot \ qx\right]S_\lambda + \left(q\cdot S\right)q_\lambda\right\}\ P_\tau \ \epsilon^{\mu\nu\lambda\tau } \label{projT} \end{eqnarray} and choosing the pertinent polarization, yields the longitudinal component \begin{eqnarray} g_L(x,Q^2)=g_1(x,Q^2)\ , \end{eqnarray} as well as the transverse combination \begin{eqnarray} g_T(x,Q^2)=g_1(x,Q^2) + g_2(x,Q^2)\ . \end{eqnarray} Also, $b=-4M\left\{\left(\frac{P\cdot q}{M}\right)^2 + 2P\cdot{qx}- \left(q\cdot{S}\right)^2\right\}$. In the Bjorken limit, which corresponds to the kinematical regime \begin{eqnarray} q_0=|\mbox{\boldmath $q$}| - M_N x \quad {\rm with}\quad |\mbox{\boldmath $q$}|\rightarrow \infty \ , \label{bjlimit} \end{eqnarray} the antisymmetric component of the hadronic tensor becomes \cite{Ja75}, \begin{eqnarray} W^{(A)}_{\mu\nu}(q)&=&\int \frac{d^4k}{(2\pi)^4} \ \epsilon_{\mu\rho\nu\sigma}\ k^\rho\ {\rm sgn}\left(k_0\right) \ \delta\left(k^2\right) \int_{-\infty}^{+\infty} dt \ {\rm e}^{i(k_0+q_0)t} \nonumber \\* && \times \int d^3x_1 \int d^3x_2 \ {\rm exp}\left[-i(\mbox{\boldmath $k$}+\mbox{\boldmath $q$})\cdot (\mbox{\boldmath $x$}_1-\mbox{\boldmath $x$}_2)\right] \nonumber \\* && \times \langle N |\left\{ {\bar \Psi}(\mbox{\boldmath $x$}_1,t){\cal Q}^2\gamma^\sigma\gamma^{5} \Psi(\mbox{\boldmath $x$}_2,0)+ {\bar \Psi}(\mbox{\boldmath $x$}_2,0){\cal Q}^2\gamma^\sigma\gamma^{5} \Psi(\mbox{\boldmath $x$}_1,t)\right\}| N \rangle \ , \label{stpnt} \end{eqnarray} where $\epsilon_{\mu\rho\nu\sigma}\gamma^\sigma \gamma^5$ is the antisymmetric combination of $\gamma_\mu\gamma_\rho\gamma_\nu$. The matrix element between the nucleon states is to be taken in the space of the collective coordinates, $A(t)$ (see eqs. (\ref{collrot}) and (\ref{nwfct})) as the object in curly brackets is an operator in this space. In deriving the expression (\ref{stpnt}) the {\it free} correlation function for the intermediate quark fields has been assumed\footnote{Adopting a dressed correlation will cause corrections starting at order twist--4 in QCD \cite{Ja96}.} after applying Wick's theorem to the product of quark currents in eq. (\ref{deften}). \cite{Ja75}. The use of the {\it free} correlation function is justified because in the Bjorken limit (\ref{bjlimit}) the intermediate quark fields carry very large momenta and are hence not sensitive to typical soliton momenta. This procedure reduces the commutator $[J_\mu(\mbox{\boldmath $x$}_1,t), J^{\dag}_\nu(\mbox{\boldmath $x$}_2,0)]$ of the quark currents in the definition (\ref{deften}) to objects which are merely bilinear in the quark fields. Consequently, in the Bjorken limit (\ref{bjlimit}) the momentum, $k$, of the intermediate quark state is highly off--shell and hence is not sensitive to momenta typical for the soliton configuration. Therefore, the use of the free correlation function is a good approximation in this kinematical regime. Accordingly, the intermediate quark states are taken to be massless, {\it cf.} eq. (\ref{stpnt}). Since the NJL model is originally defined in terms of quark degrees of freedom, quark bilinears as in eq. (\ref{stpnt}) can be computed from the functional \begin{eqnarray} \hspace{-1cm} \langle {\bar q}(x){\cal Q}^{2} q(y) \rangle&=& \int D{\bar q} Dq \ {\bar q}(x){\cal Q}^{2} q(y)\ {\rm exp}\left(i \int d^4x^\prime\ {\cal L}\right) \nonumber \\* &=&\frac{\delta}{i\delta\alpha(x,y)}\int D{\bar q} Dq \ {\rm exp}\left(i\int d^4x^\prime d^4y^\prime \left[\delta^4(x^\prime - y^\prime ){\cal L} \right. \right. \nonumber \\* && \hspace{5cm} \left. \left. +\ \alpha(x^\prime,y^\prime){\bar q}(x^\prime){\cal Q}^{2} q(y^\prime)\right] \right)\Big|_{\alpha(x,y)=0}\ . \label{gendef} \end{eqnarray} The introduction of the bilocal source $\alpha(x,y)$ facilitates the functional bosonization after which eq. (\ref{gendef}) takes the form \begin{eqnarray} \frac{\delta}{\delta\alpha(x,y)}{\rm Tr}_{\Lambda}{\rm log} \left(\delta^4(x-y)D+\alpha(x,y){\cal Q}^{2})\right) \Big|_{\alpha(x,y)=0}\ \ . \label{gendef1} \end{eqnarray} The operator $D$ is defined in eq. (\ref{dirac}). The correlation $\langle {\bar q}(x){\cal Q}^2 q(y) \rangle$ depends on the angle between $\mbox{\boldmath $x$}$ and $\mbox{\boldmath $y$}$. Since in general the functional (\ref{gendef}) involves quark states of all angular momenta ($l$) a technical difficulty arises because this angular dependence has to be treated numerically. The major purpose of the present paper is to demonstrate that polarized structure functions can indeed be computed from a chiral soliton. With this in mind we will adopt the valence quark approximation where the quark configurations in (\ref{gendef}) are restricted to the valence quark level. Accordingly the valence quark wave--function (\ref{valrot}) is substituted into eq. (\ref{stpnt}). Then only quark orbital angular momenta up to $l=2$ are relevant. From a physical point of view this approximation is justified for moderate constituent quark masses ($m\approx400{\rm MeV}$) because in that parameter region the soliton properties are dominated by their valence quark contributions \cite{Al96}, \cite{Gok96}. In particular this is the case for the axial properties of the nucleon. In the next step the polarized structure functions, $g_1(x,\mu^2)$ and $g_T(x,\mu^2)$, are extracted according to eqs. (\ref{proj1}) and (\ref{projT}). In the remainder of this section we will omit explicit reference to the scale $\mu^2$. We choose the frame such that the nucleon is polarized along the positive--$\mbox{\boldmath $z$}$ and positive--$\mbox{\boldmath $x$}$ directions in the longitudinal and transverse cases, respectively. Note also that this implies the choice ${\mbox{\boldmath $q$}}=q\hat{\mbox{\boldmath $z$}}$. When extracting the structure functions the integrals over the time coordinate in eq. (\ref{stpnt}) can readily be done yielding the conservation of energy for forward and backward moving intermediate quarks. Carrying out the integrals over $k_0$ and $k=|\mbox{\boldmath $k$}|$ gives for the structure functions \begin{eqnarray} \hspace{-1cm} g_1(x)&=&-N_C\frac{M_N}{\pi} \langle N,\frac{1}{2}\hat{\mbox{\boldmath $z$}}|\int d\Omega_{\mbox{\boldmath $k$}} k^2 \Bigg\{\tilde\psi_{\rm v}^{\dag}(\mbox{\boldmath $p$}) \left(1-\mbox{\boldmath $\alpha$}\cdot {\hat{\mbox{\boldmath $k$}}}\right)\gamma^5\Gamma \tilde\psi_{\rm v}(\mbox{\boldmath $p$}) \Big|_{k=q_0+\epsilon_{\rm v}} \nonumber \\* && \hspace{3cm} +\tilde\psi_{\rm v}^{\dag}(-\mbox{\boldmath $p$}) \left(1-\mbox{\boldmath $\alpha$}\cdot {\hat{\mbox{\boldmath $k$}}}\right)\gamma^5\Gamma \tilde\psi_{\rm v}(-\mbox{\boldmath $p$}) \Big|_{k=q_0-\epsilon_{\rm v}} \Bigg\} |N,\frac{1}{2}\hat{\mbox{\boldmath $z$}}\rangle\ , \label{valg1}\\ \hspace{-1cm} g_{T}(x)&=&g_1(x)+g_2(x) \nonumber \\* &=&-N_C\frac{M_N}{\pi} \langle N,\frac{1}{2}\hat{\mbox{\boldmath $x$}} | \int d\Omega_{\mbox{\boldmath $k$}} k^2 \Bigg\{\tilde\psi_{\rm v}^{\dag}(\mbox{\boldmath $p$}) \left(\mbox{\boldmath $\alpha$}\cdot {\hat{\mbox{\boldmath $k$}}}\right)\gamma^5\Gamma \tilde\psi_{\rm v}(\mbox{\boldmath $p$}) \Big|_{k=q_0+\epsilon_{\rm v}} \nonumber \\* && \hspace{3cm} +\tilde\psi_{\rm v}^{\dag}(-\mbox{\boldmath $p$}) \left(\mbox{\boldmath $\alpha$}\cdot {\hat{\mbox{\boldmath $k$}}}\right)\gamma^5\Gamma \tilde\psi_{\rm v}(-\mbox{\boldmath $p$}) \Big|_{k=q_0-\epsilon_{\rm v}} \Bigg\} |N,\frac{1}{2}\hat{\mbox{\boldmath $x$}} \rangle\ , \label{valgt} \end{eqnarray} where $\mbox{\boldmath $p$}=\mbox{\boldmath $k$}+\mbox{\boldmath $q$}$ and $\Gamma =\frac{5}{18}{\mbox{{\sf 1}\zr{-0.16}\rule{0.04em}{1.55ex}\zr{0.1}}} +\frac{1}{6}D_{3i}\tau_{i}$ with $D_{ij}=\frac{1}{2}\ {\rm tr}\left(\tau_{i}A(t)\tau_{j}A^{\dagger}\right)$ being the adjoint representation of the collective rotation {\it cf.} eq. (\ref{collrot}). The second entry in the states labels the spin orientation. $N_C$ appears as a multiplicative factor because the functional trace (\ref{gendef1}) includes the color trace as well. Furthermore the Fourier transform of the valence quark wave--function \begin{eqnarray} \tilde\psi_{\rm v}(\mbox{\boldmath $p$})=\int \frac{d^3x}{4\pi}\ \psi_{\rm v}(\mbox{\boldmath $x$})\ {\rm exp}\left(i\mbox{\boldmath $p$}\cdot \mbox{\boldmath $x$}\right) \label{ftval} \end{eqnarray} has been introduced. Also, note that the wave--function $\psi_{\rm v}$ contains an implicit dependence on the collective coordinates through the angular velocity $\mbox{\boldmath $\Omega$}$, {\it cf.} eq. (\ref{valrot}). The dependence of the wave--function $\tilde\psi(\pm\mbox{\boldmath $p$})$ on the integration variable ${\hat{\mbox{\boldmath $k$}}}$ is only implicit. In the Bjorken limit the integration variables may then be changed to \cite{Ja75} \begin{eqnarray} k^2 \ d\Omega_{\mbox{\boldmath $k$}} = p dp\ d\Phi\ , \qquad p=|\mbox{\boldmath $p$}|\ , \label{intdp} \end{eqnarray} where $\Phi$ denotes the azimuth--angle between $\mbox{\boldmath $q$}$ and $\mbox{\boldmath $p$}$. The lower bound for the $p$--integral is adopted when $\mbox{\boldmath $k$}$ and $\mbox{\boldmath $q$}$ are anti--parallel; $p^{\rm min}_\pm=|M_N x\mp \epsilon_{\rm v}|$ for $k=-\left(q_0\pm\epsilon_{\rm v}\right)$, respectively. Since the wave--function $\tilde\psi(\pm\mbox{\boldmath $p$})$ acquires its dominant support for $p\le M_N$ the integrand is different from zero only when $\mbox{\boldmath $q$}$ and $\mbox{\boldmath $k$}$ are anti--parallel. We may therefore take ${\hat{\mbox{\boldmath $k$}}}=-{\hat{\mbox{\boldmath $z$}}}$. This is nothing but the light--cone description for computing structure functions \cite{Ja91}. Although expected, this result is non--trivial and will only come out in models which have a current operator which, as in QCD, is formally identical to the one of non--interacting quarks. The valence quark state possesses positive parity yielding $\tilde\psi(-\mbox{\boldmath $p$})=\gamma_0 \tilde\psi(\mbox{\boldmath $p$})$. With this we arrive at the expression for the isoscalar and isovector parts of the polarized structure function in the valence quark approximation, \begin{eqnarray} \hspace{-.5cm} g^{I=0}_{1,\pm}(x)&=&-N_C\frac{5\ M_N}{18\pi} \langle N,\frac{1}{2}\hat{\mbox{\boldmath $z$}}| \int^\infty_{M_N|x_\mp|}p dp \int_0^{2\pi}d\Phi\ \nonumber \\* && \hspace{4cm}\times \tilde\psi_{\rm v}^{\dag}(\mbox{\boldmath $p$}_\mp) \left(1\pm\alpha_3\right)\gamma^5\tilde\psi_{\rm v}(\mbox{\boldmath $p$}_\mp) |N,\frac{1}{2}\hat{\mbox{\boldmath $z$}}\rangle \label{g10} \\ \hspace{-.5cm} g^{I=1}_{1,\pm}(x)&=&-N_C\frac{M_N}{6\pi} \langle N,\frac{1}{2}\hat{\mbox{\boldmath $z$}}| D_{3i} \int^\infty_{M_N|x_\mp|}p dp \int_0^{2\pi}d\Phi\ \nonumber \\* && \hspace{4cm}\times \tilde\psi_{\rm v}^{\dag}(\mbox{\boldmath $p$}_\mp)\tau_i \left(1\pm\alpha_3\right)\gamma^5\tilde\psi_{\rm v}(\mbox{\boldmath $p$}_\mp) |N,\frac{1}{2}\hat{\mbox{\boldmath $z$}}\rangle\ , \label{g11}\\ \hspace{-.5cm} g^{I=0}_{T,\pm}(x)&=&-N_C\frac{5\ M_N}{18\pi} \langle N,\frac{1}{2}\hat{\mbox{\boldmath $x$}}| \int^\infty_{M_N|x_\mp|}p dp \int_0^{2\pi}d\Phi\ \nonumber \\* && \hspace{4cm}\times \tilde\psi_{\rm v}^{\dag}(\mbox{\boldmath $p$}_\mp) \alpha_3\gamma^5\tilde\psi_{\rm v}(\mbox{\boldmath $p$}_\mp) |N,\frac{1}{2}\hat{\mbox{\boldmath $x$}}\rangle \label{gt0}\ , \\ \hspace{-.5cm} g^{I=1}_{T,\pm}(x)&=&-N_C\frac{M_N}{6\pi} \langle N,\frac{1}{2}\hat{\mbox{\boldmath $x$}}| D_{3i} \int^\infty_{M_N|x_\mp|}p dp \int_0^{2\pi}d\Phi\ \nonumber \\* && \hspace{4cm}\times \tilde\psi_{\rm v}^{\dag}(\mbox{\boldmath $p$}_\mp)\tau_i \alpha_3\gamma^5\tilde\psi_{\rm v}(\mbox{\boldmath $p$}_\mp) |N,\frac{1}{2}\hat{\mbox{\boldmath $x$}}\rangle\ , \label{gt1} \end{eqnarray} where $x_{\pm}=x\pm\epsilon_{\rm v}/{M_N}$ and ${\rm cos}(\Theta^\pm_p)={M_N}x_\pm/{p}$. The complete structure functions are given by \begin{eqnarray} g_{1}(x)&=&g^{I=0}_{1,+}(x)+g^{I=1}_{1,+}(x) -\left(g^{I=0}_{1,-}(x)-g^{I=1}_{1,-}(x)\right) \label{gone} \\* \hspace{-1cm} g_{T}(x)&=&g^{I=0}_{T,+}(x)+g^{I=1}_{T,+}(x) -\left(g^{I=0}_{T,-}(x)-g^{I=1}_{T,-}(x)\right)\ . \label{gtran} \end{eqnarray} Note also, that we have made explicit the isoscalar $\left(I=0\right)$ and isovector $\left(I=1\right)$ parts. The wave--function implicitly depends on $x$ because $\tilde\psi_{\rm v}(\mbox{\boldmath $p$}_\pm)= \tilde\psi_{\rm v}(p,\Theta^\pm_p,\Phi)$ where the polar--angle, $\Theta^\pm_p$, between $\mbox{\boldmath $p$}_\pm$ and $\mbox{\boldmath $q$}$ is fixed for a given value of the Bjorken scaling variable $x$. Turning to the evaluation of the nucleon matrix elements defined above we first note that the Fourier transform of the wave--function is easily obtained because the angular parts are tensor spherical harmonics in both coordinate and momentum spaces. Hence, only the radial part requires numerical treatment. Performing straightforwardly the azimuthal integrations in eqs. (\ref{g10}) and (\ref{g11}) reveals that the surviving isoscalar part of the longitudinal structure function, $g_{1}^{I=0}$, is linear in the angular velocity, $\mbox{\boldmath $\Omega$}$. It is this part which is associated with the proton--spin puzzle. Using the standard quantization condition, $\mbox{\boldmath $\Omega$} =\mbox{\boldmath $J$}/\ \alpha^2$, where $\alpha^2$ is the moment of inertia of the soliton and further noting that the ${\hat{\mbox{\boldmath $z$}}}$--direction is distinct, the required nucleon matrix elements are $\langle N,\frac{1}{2}\hat{\mbox{\boldmath $z$}}| J_{z}|N,\frac{1}{2}\hat{\mbox{\boldmath $z$}}\rangle=\frac{1}{2}$. Thus, $g_1^{I=0}$ is identical for all nucleon states. Choosing a symmetric ordering \cite{Al93}, \cite{Sch95} for the non--commuting operators, $D_{ia}J_j\rightarrow \frac{1}{2}\left\{D_{ia},J_j \right\}$ we find that the nucleon matrix elements associated with the cranking portion of the isovector piece, $\langle N,\pm\frac{1}{2}\hat{\mbox{\boldmath $z$}}|\left\{D_{3y},J_x \right\}|N,\pm\frac{1}{2}\hat{\mbox{\boldmath $z$}}\rangle$, vanish. With this ordering we avoid the occurance of PCAC violating pieces in the axial current. The surviving terms stem solely from the classical part of the valence quark wave--function, $\Psi_{\rm v}\left({\mbox{\boldmath $x$}}\right)$ in combination with the collective Wigner--D function, $D_{3z}$. Again singling out the ${\hat{\mbox{\boldmath $z$}}}$--direction, the nucleon matrix elements become \cite{Ad83} \begin{eqnarray} \langle N,\frac{1}{2}\hat{\mbox{\boldmath $z$}}|D_{3z}|N,\frac{1}{2}\hat{\mbox{\boldmath $z$}} \rangle = -\frac{2}{3} i_3\ , \label{matz} \end{eqnarray} where $i_3=\pm\frac{1}{2}$ is the nucleon isospin. For the transverse structure function, the surviving piece of the isoscalar contribution is again linear in the angular velocities. The transversally polarized nucleon gives rise to the matrix elements, $\langle N,\frac{1}{2}\hat{\mbox{\boldmath $x$}} |J_{x}|N,\frac{1}{2}\hat{\mbox{\boldmath $x$}}\rangle=\frac{1}{2}$. Again choosing symmetric ordering for terms arising from the cranking contribution, the nucleon matrix elements $\langle N,\frac{1}{2}\hat{\mbox{\boldmath $x$}}|\left\{D_{3y},J_y \right\}|N,\frac{1}{2}\hat{\mbox{\boldmath $x$}}\rangle$ and $\langle N,\frac{1}{2}\hat{\mbox{\boldmath $x$}}|\left\{D_{33},J_y \right\}|N,\frac{1}{2}\hat{\mbox{\boldmath $x$}}\rangle$ vanish. As in the longitudinal case, there is a surviving isovector contribution stemming solely from the classical part of the valence quark wave--function, $\Psi_{\rm v}({\mbox{\boldmath $x$}})$ in combination with the collective Wigner--D function, $D_{3x}$. Now singling out the $\hat{\mbox{\boldmath $x$}}$--direction the relevant nucleon matrix elements become \cite{Ad83}, \begin{eqnarray} \langle N,\frac{1}{2}\hat{\mbox{\boldmath $x$}}|D_{3x} |N,\frac{1}{2}\hat{\mbox{\boldmath $x$}}\rangle = -\frac{2}{3} i_3\ . \label{matx} \end{eqnarray} Explicit expressions in terms of the valence quark wave functions (\ref{gone} and \ref{gtran}) for $g^{I=0}_{1,\pm}(x)$, $g^{I=1}_{1,\pm}(x)$, $g^{I=0}_{2,\pm}(x)$ and $g^{I=1}_{,\pm}(x)$ are listed in the appendix A. Using the expressions given in the appendix A it is straightforward to verify the Bjorken sum rule \cite{Bj66} \begin{eqnarray} \Gamma_1^{p}-\Gamma_1^{n}&=&\int_{0}^{1} dx\ \left(g_{1}^{p}(x)- g_{1}^{n}(x)\right)=g_{A}/6\ , \label{bjs} \end{eqnarray} the Burkhardt--Cottingham sum rule \cite{bur70} \begin{eqnarray} \Gamma_2^{p}&=&\int_{0}^{1} dx\ g_{2}^{p}(x)=0\ , \label{bcs} \end{eqnarray} as well as the axial singlet charge \begin{eqnarray} \Gamma_1^{p}+\Gamma_1^{n}&=&\int_{0}^{1} dx\ \left(g_{1}^{p}(x)+ g_{1}^{n}(x)\right)=g_A^{0}\ , \label{gas} \end{eqnarray} in this model calculation when the moment of inertia $\alpha^2$, as well as the axial charges $g_A^0$ and $g_A$, are confined to their dominating valence quark pieces. We have used \begin{eqnarray} g_A&=&-\frac{N_C}{3}\int d^3 r {\bar\psi}_{\rm v}^{\dagger}(\mbox{\boldmath $r$})\gamma_3 \gamma_5\tau_3 \psi_{\rm v}(\mbox{\boldmath $r$}) \label{gaval} \\ g_A^0&=&\frac{N_C}{\alpha_{\rm v}^2} \int d^3 r{\bar\psi}_{\rm v}^{\dagger}(\mbox{\boldmath $r$})\gamma_3 \gamma_5\psi_{\rm v}(\mbox{\boldmath $r$}) \ . \label{ga0val} \end{eqnarray} to verify the Bjorken Sum rule as well as the axial singlet charge. This serves as an analytic check on our treatment. Here $\alpha_{\rm v}^2$ refers to the valence quark contribution to the moment of inertia, {\it i.e.} $\alpha_{\rm v}^2=(1/2)\sum_{\mu\ne{\rm v}} |\langle\mu|\tau_3|{\rm v}\rangle|^2/(\epsilon_\mu-\epsilon_{\rm v})$. The restriction to the valence quark piece is required by consistency with the Adler sum rule in the calculation of the unpolarized structure functions in this approximation \cite{wei96}. \bigskip \section{Numerical Results} \bigskip In this section we display the results of the spin--polarized structure functions calculated from eqs. (\ref{g1zro}--\ref{gton}) for constituent quark masses of $m=400{\rm MeV}$ and $450{\rm MeV}$. In addition to checking the above mentioned sum rules see eqs. (\ref{bjs})--(\ref{gas}), we have numerically calculated the first moment of $g_{1}^{p}(x,\mu^{2})$\footnote{Which in this case amounts to the Ellis--Jaffe sum rule \cite{Ja74} since we have omitted the strange degrees of freedom. A careful treatment of symmetry breaking effects indicates that the role of the strange quarks is less important than originally assumed \cite{jon90,Li95}.} \begin{eqnarray} \Gamma_1^{p}&=&\int_{0}^{1} dx\ g_{1}^{p}(x)\ , \label{ejs} \end{eqnarray} and the Efremov--Leader--Teryaev (ELT) sum rule \cite{Ef84} \begin{eqnarray} \Gamma_{\rm ETL}&=&\int_{0}^{1} dx\ x\left(g_{1}^{p}(x) +2g_{2}^{n}(x)\right)\ . \label{elts} \end{eqnarray} We summarize the results for the sum rules in table 1. When comparing these results with the experimental data one observes two short--comings, which are already known from studies of the static properties in this model. First, the axial charge $g_A\approx 0.73$ comes out too low as the experimental value is $g_A=1.25$. It has recently been speculated that a different ordering of the collective operators $D_{ai}J_j$ ({\it cf.} section 4) may fill the gap \cite{Wa93,Gok96}. However, since such an ordering unfortunately gives rise to PCAC violating contributions to the axial current \cite{Al93} and furthermore inconsistencies with $G$--parity may occur in the valence quark approximation \cite{Sch95} we will not pursue this issue any further at this time. Second, the predicted axial singlet charge $g_A^0\approx 0.6$ is approximately twice as large as the number extracted from experiment\footnote{Note that this analysis assumes $SU(3)$ flavor symmetry, which, of course, is not manifest in our two flavor model.} $0.27\pm0.04$\cite{ell96}. This can be traced back to the valence quark approximation as there are direct and indirect contributions to $g_A^0$ from both the polarized vacuum and the valence quark level. Before canonical quantization of the collective coordinates one finds a sum of valence and vacuum pieces \begin{eqnarray} g_A^0=2\left(g_{\rm v}^0+g_{\rm vac}^0\right)\Omega_3 =\frac{g_{\rm v}^0+g_{\rm vac}^0} {\alpha^2_{\rm v}+\alpha^2_{\rm vac}} \ . \label{ga0val1} \end{eqnarray} Numerically the vacuum piece is negligible, {\it i.e.} $g_{\rm vac}^0/g_{\rm v}^0\approx 2\%$. Canonical quantization subsequently involves the moment of inertia $\alpha^2=\alpha^2_{\rm v}+\alpha^2_{\rm vac}$, which also has valence and vacuum pieces. In this case, however, the vacuum part is not so small: $\alpha^2_{\rm vac}/\alpha^2\approx25\%$. Hence the full treatment of the polarized vacuum will drastically improve the agreement with the empirical value for $g_A^0$. On the other hand our model calculation nicely reproduces the Ellis--Jaffe sum rule since the empirical value is $0.136$. Note that this comparison is legitimate since neither the derivation of this sum rule nor our model imply strange quarks. While the vanishing Burkhardt--Cottingham sum rule can be shown analytically in this model, the small value for the Efremov--Leader--Teryaev sum rule is a numerical prediction. Recently, it has been demonstrated \cite{So94} that that the ELT sum rule (\ref{elts}), which is derived within the parton model, neither vanishes in the Center of Mass bag model\cite{So94} nor is supported by the SLAC E143 data \cite{slac96}. This is also the case for our NJL--model calculation as can be seen from table I. In figure 1 we display the spin structure functions $g_{1}^{p}(x,\mu^{2})$ and $g_{2}^{p}(x,\mu^{2})$ along with the twist--2 piece, $g_{2}^{WW(p)}\left(x,\mu^{2}\right)$ and twist--3 piece, ${\overline{g}}_{2}^{p}\left(x,\mu^{2}\right)$. The actual value for $\mu^2$ will be given in the proceeding section in the context of the evolution procedure. We observe that the structure functions $g_{2}^{p}(x,\mu^{2})$ and $g_{2}^{WW(p)}(x,\mu^{2})$ are well localized in the interval $0\le x\le1$, while for $g_1^{p}$ about $0.3\%$ of the first moment, $\Gamma_1^{p}=\int_{0}^{1} dx\ g_{1}^{p}(x,\mu^2)$ comes from the region, $x > 1$. The polarized structure function $g_1^{p}(x,\mu^2)$ exhibits a pronounced maximum at $x\approx0.3$ which is smeared out when the constituent quark mass increases. This can be understood as follows: In our chiral soliton model the constituent mass serves as a coupling constant of the quarks to the chiral field (see eqs. (\ref{bosact}) and (\ref{hamil})). The valence quark becomes more strongly bound as the constituent quark mass increases. In this case the lower components of the valence quark wave--function increase and relativistic effects become more important resulting in a broadening of the maximum. With regard to the Burkhardt--Cottingham sum rule the polarized structure function $g_2^{p}(x,\mu^2)$ possesses a node. Apparently this node appears at approximately the same value of the Bjorken variable $x$ as the maximum of $g_1^{p}(x,\mu^2)$. Note also that the distinct twist contributions to $g_2^{p}(x,\mu^2)$ by construction diverge as ${\rm ln}\left(x\right)$ as $x\to0$ while their sum stays finite(see section 6 for details). As the results displayed in figure 1 are the central issue of our calculation it is of great interest to compare them with the available data. As for all effective low--energy models of the nucleon, the predicted results are at a lower scale $Q^2$ than the experimental data. In order to carry out a sensible comparison either the model results have to be evolved upward or the QCD renormalization group equations have to be used to extract structure functions at a low--renormalization point. For the combination $xg_1(x)$ a parametrization of the empirical structure function is available at a low scale \cite{Gl95}\footnote{These authors also provide a low scale parametrization of quark distribution functions. However, these refer to the distributions of perturbatively interacting partons. Distributions for the NJL--model constituent quarks could in principle be extracted from eqs. (\ref{g10})--(\ref{gt1}). It is important to stress that these distributions may not be compared to those of ref \cite{Gl95} because the associated quarks fields are different in nature.}. In that study the experimental high $Q^2$ data are evolved to the low--renormalization point $\mu^2$, which is defined as the lowest $Q^2$ satisfying the positivity constraint between the polarized and unpolarized structure functions. In a next--to--leading order calculation those authors found $\mu^2=0.34{\rm GeV}^2$ \cite{Gl95}. In figure 2 we compare our results for two different constituent quark masses with that parametrization. We observe that our predictions reproduce gross features like the position of the maximum. This agreement is the more pronounced the lower the constituent quark is, {\it i.e.} the agreement improves as the applicability of the valence quark approximation becomes more justified. Unfortunately, such a parametrization is currently not available for the transverse structure function $g_T(x)$ (or $g_2(x)$). In order to nevertheless be able to compare our corresponding results with the (few) available data we will apply leading order evolution techniques to the structure functions calculated in the valence quark approximation to the NJL--soliton model. This will be subject of the following section. \bigskip \section{Projection and Evolution} \bigskip One notices that our baryon states are not momentum eigenstates causing the structure functions (see figures 1 and 2) not to vanish exactly for $x>1$ although the contributions for $x>1$ are very small. This short--coming is due to the localized field configuration and thus the nucleon not being a representation of the Poincar\'{e} group which is common to the low--energy effective models. The most feasible procedure to cure this problem is to apply Jaffe's prescription \cite{Ja80}, \begin{eqnarray} f(x)\longrightarrow \tilde f(x)= \frac{1}{1-x}f\left(-{\rm log}(1-x)\right) \label{proj} \end{eqnarray} to project any structure function $f(x)$ onto the interval $[0,1]$. In view of the kinematic regime of DIS this prescription, which was derived in a Lorentz invariant fashion within the 1+1 dimensional bag model, is a reasonable approximation. It is important to note in the NJL model the unprojected nucleon wave--function (including the cranking piece\footnote{Which in fact yields the leading order to the Adler sum rule, $F_1^{\nu p}\ - F_1^{{\bar \nu}p}$ \cite{wei96} rather than being a correction.}, see \ref{valrot}) is anything but a product of Dirac--spinors. In this context, techniques such as Peierls--Yoccoz\cite{Pei57} (which does not completely enforce proper support \cite{Sig90}, $0\le x\le1$ nor restore Lorentz invariance, see \cite{Ard93}) appear to be infeasible. Thus, given the manner in which the nucleon arises in chiral--soliton models Jaffe's projection technique is quite well suited. It is also important to note that, by construction, sum rules are not effected by this projection, {\it i.e.} $\int_0^\infty dxf(x)= \int_0^1 dx \tilde f(x)$. Accordingly the sum--rules of the previous section remain intact. With regard to evolution of the spin--polarized structure functions applying the OPE analysis of Section 2, Jaffe and Ji brought to light that to leading order in $1/Q^{2}$, $g_1(x,Q^2)$ receives only a leading order twist--2 contribution, while $g_2(x,Q^2)$ possesses contributions from both twist--2 and twist--3 operators; the twist--3 portion coming from spin--dependent gluonic--quark correlations \cite{Ja90},\cite{Ji90} (see also, \cite{ko79} and \cite{sh82}). In the {\em impulse approximation} \cite{Ja90}, \cite{Ji90} these leading contributions are given by \begin{eqnarray} \hspace{-2cm} \lim_{Q^2\to\infty} \int_{0}^{1} dx\ x^{n} g_{1}(x,Q^2)&=&\frac{1}{2}\sum _{i}\ {\cal O}_{2,i}^{n}\ \ ,\ \ n=0,2,4,\ldots\ , \label{ltc1} \\ \lim_{Q^2\to\infty} \int_{0}^{1} dx\ x^{n}\ g_{2}(x,Q^2)&=&-\frac{n}{2\ (n+1)} \sum_{i} \left\{ {\cal O}_{2,i}^{n} -{\cal O}_{3,i}^{n} \right\},\ n=2,4,\ldots\ . \label{ltc2} \end{eqnarray} Note that there is no sum rule for the first moment, $\Gamma_{2}(Q^2)=\int_{0}^{1}\ dx g_{2}(x,Q^2)$, \cite{Ja90}. Sometime ago Wandzura and Wilczek \cite{wan77} proposed that $g_2(x,Q^2)$ was given in terms of $g_1(x,Q^2)$, \begin{eqnarray} g_{2}^{WW}(x,Q^2)=-\ g_{1}(x,Q^2)+\ \int_{x}^{1}\frac{dy}{y}\ g_{1}(y,Q^2) \label{ww} \end{eqnarray} which follows immediately from eqs. (\ref{ltc1}) and (\ref{ltc2}) by neglecting the twist--3 portion in the sum in (\ref{ltc2}). One may reformulate this argument to extract the twist--3 piece \begin{eqnarray} {\overline{g}}_{2}(x,Q^2)\ =\ g_{2}(x,Q^2)\ -\ g_{2}^{WW}(x,Q^2)\ , \end{eqnarray} since, \begin{eqnarray} \int_{0}^{1} dx\ x^{n}\ {\overline{g}}_{2}(x,Q^2)=\frac{n}{2\ (n+1)} \sum_{i} {\cal O}_{3,i}^{n}\ \ , \ n=2,4,\ \ldots \ . \end{eqnarray} In the NJL model as in the bag--model there are no explicit gluon degrees of freedom, however, in both models twist--3 contributions to $g_2(x,\mu^2)$ exist. In contrast to the bag model where the bag boundary simulates the quark--gluon and gluon--gluon correlations \cite{So94} in the NJL model the gluon degrees of freedom, having been ``integrated" out, leave correlations characterized by the four--point quark coupling $G_{\rm NJL}$. This is the source of the twist--3 contribution to $g_2(x,\mu^2)$, which is shown in figure 1. For $g_{1}\left(x,Q^2\right)$ and the twist--2 piece $g_2^{WW}\left(x,Q^2\right)$ we apply the leading order (in $\alpha_{QCD}(Q^2)$) Altarelli--Parisi equations \cite{Al73} to evolve the structure functions from the model scale, $\mu^2$, to that of the experiment $Q^2$, by iterating \begin{eqnarray} g(x,t+\delta{t})=g(x,t)\ +\ \delta t\frac{dg(x,t)}{dt}\ , \end{eqnarray} where $t={\rm log}\left(Q^2/\Lambda_{QCD}^2\right)$. The explicit expression for the evolution differential equation is given by the convolution integral, \begin{eqnarray} \frac{d g(x,t)}{dt}&=&\frac{\alpha(t)}{2\pi} g(x,t)\otimes P_{qq}(x) \nonumber \\* \hspace{1cm} &=&\frac{\alpha(t)}{2\pi} C_{R}(F)\int^1_{x}\ \frac{dy}{y}P_{qq}\left(y\right) g\left(\frac{x}{y},t\right) \label{convl} \end{eqnarray} where the quantity $P_{qq}\left(z\right)=\left(\frac{1+z^2}{1-z^2}\right)_{+}$ represents the quark probability to emit a gluon such that the momentum of the quark is reduced by the fraction $z$. $C_{R}(f)=\frac{{n_{f}^{2}}-1}{2{n_{f}}}$ for $n_f$--flavors, $\alpha_{QCD}=\frac{4\pi}{\beta\log\left(Q^2/ \Lambda^2\right)}$ and $\beta=(11-\frac{2}{3}n_f)$. Employing the ``+" prescription\cite{Al94} yields \begin{eqnarray} \frac{d\ g(x,t)}{dt}&=&\frac{2C_{R}(f)}{9\ t} \left\{\ \left(x + \frac{x^{2}}{2}+2\log(1-x)\right)g(x,t) \right. \nonumber \\*&& \hspace{1cm} \left. +\ \int^{1}_{x}\ dy \left(\frac{1+y^2}{1-y}\right) \left[\frac{1}{y}\ g\left(\frac{x}{y},t\right)-g(x,t)\right]\ \right\}\ . \label{evol} \end{eqnarray} As discussed in section 2 the initial value for integrating the differential equation is given by the scale $\mu^2$ at which the model is defined. It should be emphasized that this scale essentially is a new parameter of the model. For a given constituent quark mass we fit $\mu^2$ to maximize the agreement of the predictions with the experimental data on previously \cite{wei96} calculated unpolarized structure functions for (anti)neutrino--proton scattering: $F_2^{\nu p}-F_2^{\overline{\nu} p}$. For the constituent quark mass $m=400{\rm MeV}$ we have obtained $\mu^2\approx0.4{\rm GeV}^2$. One certainly wonders whether for such a low scale the restriction to first order in $\alpha_{QCD}$ is reliable. There are two answers. First, the studies in this section aim at showing that the required evolution indeed improves the agreement with the experimental data and, second, in the bag model it has recently been shown \cite{St95} that a second order evolution just increases $\mu^2$ without significantly changing the evolved data. In figure 3 we compare the unevolved, projected, structure function $g_1^{p}\left(x,\mu^{2}\right)$ with the one evolved from $\mu^{2}=0.4{\rm GeV}^2$ to $Q^2=3.0{\rm GeV}^2$. Also the data from the E143-collaboration from SLAC\cite{slac95a} are given. Furthermore in figure 3 we compare the projected, unevolved structure function $g_2^{WW(p)}\left(x,\mu^{2}\right)$ as well as the one evolved to $Q^2=5.0{\rm GeV}^2$ with the data from the recent E143-collaboration at SLAC\cite{slac96}. As expected we observe that the evolution pronounces the structure function at low $x$; thereby improving the agreement with the experimental data. This change towards small $x$ is a general feature of the projection and evolution process and presumably not very sensitive to the prescription applied here. In particular, choosing an alternative projection technique may easily be compensated by an appropriate variation of the scale $\mu^2$. While the evolution of the structure function $g_{1}\left(x,Q^2\right)$ and the twist--2 piece $g_2^{WW}\left(x,Q^2\right)$ from $\mu^2$ to $Q^2$ can be performed straightforwardly using the ordinary Altarelli--Parisi equations this is not the case with the twist--3 piece ${\overline{g}}_{2}(x,Q^2)$. As the twist--3 quark and quark--gluon operators mix the number of independent operators contributing to the twist--3 piece increases with $n$, where $n$ refers to the $n^{\underline{\rm th}}$ moment\cite{sh82}. We apply an approximation (see appendix B) suggested in\cite{Ali91} where it is demonstrated that in $N_c\to \infty$ limit the quark operators of twist--3 decouple from the evolution equation for the quark--gluon operators of the same twist resulting in a unique evolution scheme. This scheme is particularly suited for the NJL--chiral soliton model, as the soliton picture for baryons is based on $N_c\rightarrow \infty$ arguments\footnote{This scheme has also employed by Song\cite{So94} in the Center of Mass bag model.}. In figure 4 we compare the projected unevolved structure function ${\overline{g}}_{2}^{p}(x,\mu^2)$ evolved to $Q^2=5.0{\rm GeV}^2$ using the scheme suggested in \cite{Ali91}. In addition we reconstruct $g_2^{p}\left(x,Q^2\right)$ at $Q^2=3.0{\rm GeV}^2$ from $g_2^{WW(p)}\left(x,Q^2\right)$ and ${\overline{g}}_{2}(x,Q^2)$ and compare it with the recent SLAC data\cite{slac96} for $g_2^{p}\left(x,Q^2\right)$. As is evident our model calculation of $g_2^{p}\left(x,Q^2\right)$, built up from its twist--2 and twist--3 pieces, agrees reasonably well with the data although the experimental errors are quite large. \bigskip \section{Summary and Outlook} \bigskip In this paper we have presented the calculation of the polarized nucleon structure functions $g_1\left(x,Q^2\right)$ and $g_2\left(x,Q^2\right)$ within a model which is based on chiral symmetry and its spontaneous breaking. Specifically we have employed the NJL chiral soliton model which reasonably describes the static properties of the nucleon \cite{Al96}, \cite{Gok96}. In this model the current operator is formally identical to the one in an non--interacting relativistic quark model. While the quark fields become functionals of the chiral soliton upon bosonization, this feature enables one calculate the hadronic tensor. From this hadronic tensor we have then extracted the polarized structure functions within the valence quark approximation. As the explicit occupation of the valence quark level yields the major contribution (about 90\%) to the associated static quantities like the axial charge this presumably is a justified approximation. When cranking corrections are included this share may be reduced depending on whether or not the full moment of inertia is substituted. It needs to be stressed that in contrast to {\it e.g.} bag models the nucleon wave--function arises as a collective excitation of a non--perturbative meson field configuration. In particular, the incorporation of chiral symmetry leads to the distinct feature that the pion field cannot be treated perturbatively. Because of the hedgehog structure of this field one starts with grand spin symmetric quark wave--functions rather than direct products of spatial-- and isospinors as in the bag model. On top of these grand spin wave--functions one has to include cranking corrections to generate states with the correct nucleon quantum numbers. Not only are these corrections sizable but even more importantly one would not be able to make any prediction on the flavor singlet combination of the polarized structure functions without them. The structure functions obtained in this manner are, of course, characterized by the scale of the low--energy effective model. We have confirmed this issue by obtaining a reasonable agreement of the model predictions for the structure function $g_1$ of the proton with the low--renormalization point parametrization of ref \cite{Gl95}. In general this scale of the effective model essentially represents an intrinsic parameter of a model. For the NJL--soliton model we have previously determined this parameter from the behavior of the unpolarized structure functions under the Altarelli--Parisi evolution \cite{wei96}. Applying the same procedure to the polarized structure functions calculated in the NJL model yields good agreement with the data extracted from experiment, although the error bars on $g_1\left(x,Q^2\right)$ are still sizable. In particular, the good agreement at low $x$ indicates that to some extend gluonic effects are already incorporated in the model. This can be understood by noting that the quark fields, which enter our calculation, are constituent quarks. They differ from the current quarks by a mesonic cloud which contains gluonic components. Furthermore, the existence of gluonic effects in the model would not be astonishing because we had already observed from the non--vanishing twist--3 part of $g_2\left(x,Q^2\right)$, which in the OPE is associated with the quark--gluon interaction, that the model contains the main features allocated to the gluons. There is a wide avenue for further studies in this model. Of course, one would like to incorporate the effects of the polarized vacuum, although one expects from the results on the static axial properties that their direct contributions are negligible. It may be more illuminating to include the strange quarks within the valence quark approximation. This extension of the model seems to be demanded by the analysis of the proton spin puzzle. Technically two changes will occur. First, the collective matrix elements will be more complicated than in eqs. (\ref{matz}) and (\ref{matx}) because the nucleon wave--functions will be distorted $SU(3)$ $D$--functions in the presence of flavor symmetry breaking \cite{Ya88,wei96a}. Furthermore the valence quark wave--function (\ref{valrot}) will contain an additional correction due to different non--strange and strange constituent quark masses \cite{We92}. When these corrections are included direct information will be obtained on the contributions of the strange quarks to polarized nucleon structure functions. In particular the previously developed generalization to three flavors \cite{We92} allows one to consistently include the effects of flavor symmetry breaking. \bigskip \acknowledgements This work is supported in part by the Deutsche Forschungsgemeinschaft (DFG) under contract Re 856/2-2. LG is grateful for helpful comments by G. R. Goldstein. \bigskip \bigskip
proofpile-arXiv_065-482
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} Instantons~\cite{bpst} are well known to represent tunnelling transitions in non-abelian gauge theories between degenerate vacua of different topology. These transitions induce processes which are {\it forbidden} in perturbation theory, but have to exist in general~\cite{th} due to Adler-Bell-Jackiw anomalies. Correspondingly, these processes imply a violation of certain fermionic quantum numbers, notably, $B+L$ in the electro-weak gauge theory and chirality ($Q_5$) in (massless) QCD. An experimental discovery of such a novel, non-perturbative manifestation of non-abelian gauge theories would clearly be of basic significance. A number of results has revived the interest in instanton-induced processes during recent years: \begin{itemize} \item First of all, it was shown~\cite{r,m} that the generic exponential suppression of these tunnelling rates, $\propto \exp (-4\pi /\alpha )$, may be overcome at {\it high energies}, mainly due to multi-gauge boson emission in addition to the minimally required fermionic final state. \item A pioneering and encouraging theoretical estimate of the size of the instanton ($I$) induced contribution to the gluon structure functions in deep-inelastic scattering was recently presented in Ref.~\cite{bb}. The summation over the $I$-induced multi-particle final state was implicitly performed by starting from the optical theorem for the virtual $\gamma^\ast g\rightarrow \gamma^\ast g $ forward amplitude. The strategy was then to evaluate the contribution to the functional integral coming from the vicinity of the instanton-antiinstanton ($\iai$) configuration in Euclidean space, to analytically continue the result to Minkowski space and, finally, to take the imaginary part. While the instanton-induced contribution to the gluon structure functions turned out to be small at larger values of the Bjorken variable $x$, it was found in Ref.~\cite{bb} to increase dramatically towards smaller $x$. \item Last not least, a systematic phenomenological and theoretical study is under way~\cite{rs,grs,rs1,ggmrs}, which clearly indicates that deep-inelastic $e p$ scattering at HERA now offers a unique window to experimentally detect QCD-instanton induced processes through their characteristic final-state signature. The searches for instanton-induced events have just started at HERA and a first upper limit of $0.9$ nb at $95\%$ confidence level for the cross-section of QCD-instanton induced events has been placed by the H1 Collaboration~\cite{H1}. New, improved search strategies are being developped~\cite{ggmrs} with the help of a Monte Carlo generator (QCDINS 1.3)~\cite{grs} for instanton-induced events. \end{itemize} The central question is, of course, whether instanton-induced processes in deep-inelastic scattering can both be reliably computed and experimentally measured. In particular, whether contributions associated with the non-perturbative vacuum structure can be controlled in the same way as perturbative short-distance corrections, in terms of a hard scale ${\mathcal Q}$. In the work of Refs.~\cite{bb,bbgg} on deep-inelastic scattering, the integrals over the instanton size $\rho$ were found to be infrared (IR) divergent, like in a number of previous instanton calculations in different areas. Yet, the authors claimed that this problem does not affect the possibility, to isolate in deep-inelastic scattering a well-defined, IR-finite and sizable instanton contribution in the regime of small QCD-gauge coupling, on account of the (large) photon virtuality $Q^2$. The IR-divergent pieces of the $I$-size integrals were supposed to be factorizable into the parton distributions, which anyway have to be extracted from experiment at some reference scale. On the other hand, also IR-finite instanton contributions to certain observables in momentum space have been found in the past~\cite{as,early}. In this ideal case, the size of the contributing instantons is limited by the inverse momentum scale ${\mathcal Q}^{-1}$ of the experimental probe, as one might intuitively expect. No {\it ad hoc} cutoff or assumption about the behaviour of large, overlapping instantons need be introduced. A main issue of the present work is to shed further light on these important questions around the IR-behaviour associated with the instanton size in deep-inelastic scattering. This paper represents the first of several papers in preparation~\cite{mrs}, containing our theoretical results on $I$-induced processes in the deep-inelastic regime. \begin{figure} \begin{center} \epsfig{file=ampl_gen.ps,width=13cm} \caption[dum]{\label{f1} Instanton-induced {\it chirality-violating} process,\\ $\gamma^{\ast} +{\rm g}\rightarrow \sum_{\rm flavours}^{n_{f}}\left[ \overline{{\rm q}_{L}} +{\rm q}_{R}\right] +n_{g}\,{\rm g}$, corresponding to three massless flavours ($n_{f}=3$).} \end{center} \end{figure} For clarity, let us reduce here the realistic task of evaluating the $I$-induced cross-sections of the chirality violating multi-particle processes (illustrated in Fig.~\ref{f1}) \begin{equation} \gamma^{\ast}+{\rm g}\Rightarrow \overline{{\rm u}_{L}}+{\rm u}_{R}\, + \overline{{\rm d}_{L}}+{\rm d}_{R}\, +\overline{{\rm s}_{L}}+{\rm s}_{R}\, +n_g\, {\rm g}, \end{equation} to the detailed study of the {\it simplest} one, without additional gluons and with just one massless flavour ($n_f=1$), \begin{equation} \gamma^{\ast}(q)+{\rm g}(p)\Rightarrow \overline{{\rm q}_{L}}(k_{1})+{\rm q}_{R}(k_{2})\, . \end{equation} The price is, of course, that this process only represents a small fraction of the total $I$-induced contribution to the gluon structure functions. However, there are also a number of important virtues: The present calculation provides a clean and explicit discussion of most of the crucial steps involved in our subsequent task~\cite{mrs} to calculate the {\it dominant} $I$-induced contributions. While unessential technical complications have been eliminated here, a generalization to the realistic case with gluons and more flavours in the final state is entirely straightforward~\cite{mrs,q96}. We shall explicitly calculate the corresponding fixed angle cross-section and the contributions to the gluon structure functions in leading semi-classical approximation within standard instanton perturbation theory (Sect.~3). Gauge invariance is kept manifest along the calculation and we may compare at various stages with the appropriate chirality-conserving process, calculated in leading order of perturbative QCD. As a central result of this paper and unlike Ref.~\cite{bb}, we find {\it no} IR divergencies associated with the integration over the instanton size $\rho$, which can even be perfomed analytically. We are able to demonstrate explicitly that the typical hard scale ${\mathcal Q}$ in deep-inelastic scattering provides a {\it dynamical} infrared cutoff for the instanton size, $\rho \mbox{\,\raisebox{.3ex {\mathcal O}(1/{\mathcal Q})$. Additional gluons in the final state will not change this conclusion, as is briefly outlined in Sect.~4. Thus, deep-inelastic scattering may indeed be viewed as a distinguished process for studying manifestations of QCD-instantons. \section{Setting the Stage} Let us start with the matrix elements ${\mathcal T}^\mu (q,p; k_{1},\ldots ,k_{n})$ of the general, exclusive photon-parton reactions \begin{equation} \gamma^{\ast}(q)+ p\rightarrow k_{1}+\ldots +k_{n}\, , \end{equation} in terms of which we form the inclusive structure tensor ${\mathcal W}^{\mu \nu}_{p}$ of the parton $p$, \begin{eqnarray} {\mathcal W}^{\mu \nu}_{ p} (q,p ) & =& \sum_{n=1}^\infty {\mathcal W}^{\mu \nu\, (n)}_{ p} (q,p ) \, , \label{wmunu} \hspace{24pt} {\rm with} \\[10pt] {\mathcal W}^{\mu \nu (n)}_{ p} ( q,p )& =& \frac{1}{4\,\pi}\, \int dPS^{(n)}\, {\mathcal T}^\mu (q,p;k_{1},\ldots,k_{n}) {\mathcal T}^{\nu\,\ast}(q,p;k_{1},\ldots ,k_{n} ) , \label{wmunu_n} \end{eqnarray} and \begin{eqnarray} \int dPS^{(n)} = \prod_{j=1}^{n} \int \frac{d^{4}k_{j}}{(2\,\pi)^{3}}\, \delta^{(+)}\left( k_{j}^{2}\right) (2\,\pi)^{4}\, \delta^{(4)}\left( q+p-k_{1}-\ldots -k_{n}\right) \, . \end{eqnarray} Averaging over colour and spin of the initial state is implicitly understood in Eq.~(\ref{wmunu_n}); the index $n$ is to label besides the final state partons also their spin and colour degrees of freedom. For exclusive $2\rightarrow 2$ processes, $\gamma^{\ast}(q) +p\rightarrow k_{1}+k_{2}$, the differential cross-section is then expressed as \begin{equation} \label{diffcross} \frac{d\sigma}{dt}= 8\,\pi^{2}\,\alpha\ \frac{x}{Q^{2}}\, \left[ -g_{\mu \nu}\, \frac{d{\mathcal W}^{\mu \nu\, (2)}_{p}}{dt} \right] \, , \end{equation} where $Q^{2}=-q^{2}$ denotes the photon virtuality and \begin{equation} \label{t} t=\left( q-k_{1}\right)^{2}=\left( p-k_{2}\right)^{2} \, , \end{equation} the momentum transfer squared. In general, each final state, $k_{1}+\ldots +k_{n}$, contributes to the structure functions ${\mathcal F}_{i\,{ p}}$ of the parton $p$ via the projections \begin{eqnarray} {\mathcal F}^{(n)}_{2\,{ p}} (x ,Q^2) &= & \left[- g_{\nu\mu }\, +6\,x\,\frac{p_\mu p_\nu}{p\cdot q} \right] x\,{\mathcal W}_{ p}^{\mu \nu\, (n)} (q,p) \, , \label{projF2} \\[10pt] {\mathcal F}^{(n)}_{L\,{ p}} (x ,Q^2) &= & 4\,x^2\,\frac{p_\mu p_\nu}{p\cdot q} {\mathcal W}_{ p}^{\mu \nu\, (n)} (q,p) \ , \label{projFL} \hspace{24pt} {\rm such\ that} \\[10pt] {\mathcal F}_{i\,{ p}}(x,Q^{2})&=&\sum_{n=1}^{\infty} {\mathcal F}^{(n)}_{i\,{ p}}(x,Q^{2})\, , \end{eqnarray} with \begin{equation} x\equiv \frac{Q^2}{2\,p\cdot q} \, . \end{equation} denoting the Bjorken variable of the photon-parton subprocess. The spin averaged proton structure functions $F_2$ and $F_L$, appearing (in the one photon exchange approximation) in the unpolarized inclusive lepto-production cross-section as \begin{equation} \frac{d^2\sigma}{dx_{\rm Bj}\, d\ybj } = \frac{4\pi\alpha^2}{Sx_{\rm Bj}^2\ybj^2} \left[ \left\{ 1-\ybj +\frac{\ybj^2}{2} \right\} F_2 (x_{\rm Bj} ,Q^2) - \frac{\ybj^2}{2} F_L (x_{\rm Bj} ,Q^2 ) \right] , \end{equation} are expressed via a standard convolution in terms of the parton structure functions ${\mathcal F}_{i\,{p}}$ and corresponding parton densities $f_{ p}$, \begin{equation} F_i\, (x_{\rm Bj} ,Q^2) = \sum_{p=q,g} \int_{x_{\rm Bj}}^1 \frac{dx}{x}\, f_{ p}\left( \frac{x_{\rm Bj}}{x}\right) \, \frac{x_{\rm Bj}}{x}\, {\mathcal F}_{i\,{ p}} (x,Q^2) \ ,\ \ i=2,L\, . \label{protonstruc} \end{equation} Here, $\sqrt{S}$ is the center-of-mass (c.m.) energy of the lepton-hadron system. The corresponding Bjorken variables are defined as usual \begin{equation} x_{\rm Bj} \equiv \frac{Q^2}{2\,P\cdot q} ; \hspace{48pt} \ybj \equiv \frac{P\cdot q}{P\cdot k} , \end{equation} where $P\,(k)$ is the four-momentum of the incoming proton (lepton). \begin{figure} \begin{center} \epsfig{file=ampl_pt.ps,width=9cm} \caption[dum]{\label{f2} Perturbative {\it chirality-conserving} process, $\gamma^{\ast}(q)+{\rm g}(p )\rightarrow \overline{{\rm q}_{L}}(k_{1})+{\rm q}_{L}(k_{2})$. } \end{center} \end{figure} As outlined in the Introduction, we shall consider in this paper only the contributions from the simplest instanton-induced, {\it chirality-violating} photon-gluon process, corresponding to one massless quark flavour $(n_{f}=1)$, \begin{equation} \gamma^{\ast}(q)+{\rm g}(p)\rightarrow \overline{{\rm q}_{L}}(k_{1})+{\rm q}_{R}(k_{2})\ ; \hspace{24pt} \left( \triangle Q_{5}\equiv \triangle \left( Q_{R}-Q_{L}\right)=2\right) \, . \label{instproc} \end{equation} It will be very instructive to compare with the appropriate leading-order perturbative QCD amplitudes for the {\it chirality-conserving} process (Fig.~\ref{f2}), \begin{equation} \gamma^{\ast}(q)+{\rm g}(p)\rightarrow \overline{{\rm q}_{L}}(k_{1})+{\rm q}_{L}(k_{2})\ ; \hspace{24pt} \left( \triangle Q_{5}\equiv \triangle \left( Q_{R}-Q_{L}\right)=0\right) \, . \label{pertproc} \end{equation} at the various stages of the instanton calculation. Therefore, for reference, let us summarize the well-known perturbative results next (see any textbook on perturbative QCD, e.g. Ref.~\cite{f}). Let us use two-component Weyl-notation for the (massless) fermions involved, in order to facilitate the comparison with the instanton calculation later on. The leading-order amplitude for the perturbative process (\ref{pertproc}) (Fig.~\ref{f2}) then reads, \begin{eqnarray} \label{ampl_pt} \lefteqn{{\mathcal T}^{A}_{\mu \,\mu ^{\prime}} \left( \gamma^{\ast}+{\rm g}\rightarrow \overline{{\rm q}_{L}}+{\rm q}_{L}\right) =} \\[0.5ex] \nonumber && e_{q}\,g_s\,t^{A}\, \chi_{L}^{\dagger}(k_{2}) \left[ \overline{\sigma}_{\mu ^{\prime}}\frac{(q-k_{1})}{(q-k_{1})^{2}} \overline{\sigma}_\mu -\overline{\sigma}_\mu \frac{(q-k_{2})}{(q-k_{2})^{2}} \overline{\sigma}_{\mu ^{\prime}} \right] \chi_{L}(k_{1})\ , \end{eqnarray} where the two-component Weyl-spinors $\chi_{L,\,R}$ satisfy the Weyl-equations, \begin{equation} \overline{k}\,\chi_L(k)=0 \, ;\hspace{24pt} k\,\chi_R(k) =0 \, , \label{weyleq} \end{equation} and \begin{equation} \chi_L(k)\,\chi_L^\dagger (k) = k \, ; \hspace{24pt} \chi_R(k)\,\chi_R^\dagger (k) = \overline{k} \, . \label{compl} \end{equation} In Eqs.~(\ref{ampl_pt}-\ref{compl}) and throughout the paper we use the abbreviations, \begin{equation} v \equiv v_\mu \,\sigma^\mu \, ; \hspace{24pt} \overline{v} \equiv v_\mu \,\overline{\sigma}^\mu \, ,\ \mbox{for any four-vector $v_\mu $}, \end{equation} where the familiar $\sigma$-matrices\footnote{\label{sigmas} We use the standard notations, in Minkowski space: $\sigma_\mu =(1,\vec{\sigma})$, $\overline{\sigma}_\mu =(1,-\vec{\sigma})$, and in Euclidean space: $\sigma_\mu =(-\ii\,\vec{\sigma},1)$, $\overline{\sigma}_\mu = (\ii\,\vec{\sigma},1)$, where $\vec{\sigma}$ are the Pauli matrices.} satisfy, \begin{equation} \sigma_\mu \overline{\sigma}_\nu +\sigma_\nu \overline{\sigma}_\mu = 2\,g_{\mu \nu} \, . \label{com} \end{equation} Finally, in Eq.~(\ref{ampl_pt}), $t^A,A=1,\ldots ,8$, are the SU(3) generators, $e_q$ is the quark charge in units of the electric charge $e$, and $g_s$ is the SU(3) gauge coupling. With help of Eqs.~(\ref{weyleq}), (\ref{com}) and the on-shell conditions $k_1^2=k_2^2=0$, the gauge-invariance constraints, \begin{equation} q^\mu \,{\mathcal T}^{A}_{\mu \,\mu ^{\prime}}=0\, ;\hspace{24pt} {\mathcal T}^{A}_{\mu \,\mu ^{\prime}}\,p^{ \mu ^{\prime}}=0\, , \label{gaugeinv} \end{equation} are easily checked. Next, we obtain the leading-order contribution of the process (\ref{pertproc}) to ${\mathcal W}^{\mu \nu\,(2)}_{g}(q,p)$ by contracting Eq.~(\ref{ampl_pt}) with the gluon polarization vector $\epsilon_{g}^{\mu ^{\prime}}(p)$ and taking the traces in Eq.~(\ref{wmunu_n}) by means of relations (\ref{compl}) and (\ref{com}). Averaging over the initial-state gluon polarization and colour amounts to an overall factor $1/16$. The final result for the projections needed in Eqs.~(\ref{diffcross}), (\ref{projF2}), and (\ref{projFL}) then reads \begin{eqnarray} \label{gmunupt} \lefteqn{ -g_{\mu \nu}\, \frac{d{\mathcal W}^{\mu \nu\,{(2)}}_{g}}{dt} \,\left( \gamma^{\ast}+{\rm g}\rightarrow \overline{{\rm q}_{L}}+{\rm q}_{L}\right) =} \\[0.5ex] \nonumber && e_q^2\, \frac{\alpha_s}{4\,\pi}\, \frac{x}{2\,Q^2}\, \left[ \frac{u}{t}+\frac{t}{u}-2\,\frac{1-x}{x}\,\frac{Q^4}{tu} \right] \, , \end{eqnarray} \begin{equation} \label{pmupnupt} \frac{p_\mu \,p_{\nu}}{p\cdot q} \, \frac{d{\mathcal W}^{\mu \nu\,{(2)}}_{g}}{dt} \,\left( \gamma^{\ast}+{\rm g}\rightarrow \overline{{\rm q}_{L}}+{\rm q}_{L}\right) = e_q^2\,\frac{\alpha_s}{4\,\pi}\, \frac{x\,(1-x)}{Q^2} \, , \end{equation} where $u=-t-Q^2/x$. Upon integrating Eq.~(\ref{gmunupt}) over $t$, we encounter the familiar collinear divergencies for $t,u\rightarrow 0$. In order to isolate the hard contributions to the gluon structure functions, it is adequate to regularize the collinear singularities by introducing an infrared cutoff scale $\mu _{c}$ in the integration limits, $\{-Q^{2}/x+\mu _{c}^{2},-\mu _{c}^{2}\}$. On account of Eqs.~(\ref{projF2}), (\ref{projFL}), one then obtains the familiar results\footnote{\label{foot1}When comparing with the literature, one has to remember that we considered only the production of a $\overline{{\rm q}_{L}}\,{\rm q}_{L}$ pair (c.\,f. Eq.~(\ref{pertproc})). In the full ${\mathcal O}(\alpha_{s})$ contribution to the gluon structure functions, the production of a $\overline{{\rm q}_{R}}\,{\rm q}_{R}$ pair has also to be included. This amounts to multiplying Eqs.~(\ref{F2pert}), (\ref{FLpert}), by a factor of 2.} in the Bjorken limit, \begin{eqnarray} \label{F2pert} && {\mathcal F}_{2\,g}^{(\overline{{\rm q}_L}{\rm q}_L)} \, (x,Q^2;\mu _{c}^{2})= e_q^2\, \frac{\alpha_s}{2\,\pi}\,\times \\[0.8ex] \nonumber && x\, \left[ P_{qg}(x)\, \ln\left(\frac{Q^{2}}{\mu _{c}^{2}}\right) +P_{qg}(x)\,\ln \left(\frac{1}{x}\right) -\frac{1}{2} +3\,x\,(1-x) \right] \left[1+{\mathcal O}\left(\frac{\mu _{c}^{2}}{Q^{2}}\right)\right] \, , \end{eqnarray} \begin{equation} {\mathcal F}_{L\,g}^{(\overline{{\rm q}_L}{\rm q}_L)}\,(x,Q^2)= e_q^2\,\frac{\alpha_s}{\pi}\, x^{2}\,(1-x) \, , \label{FLpert} \end{equation} with the splitting function \begin{equation} \label{pqg} P_{qg}(x) \equiv \frac{1}{2}\,\left(x^{2}+(1-x)^{2}\right)\, . \end{equation} Of course, the {\it finite} part of the structure function ${\mathcal F}_{2\,g}^{(\overline{{\rm q}_L}{\rm q}_L)}$, that is everything except for the large logarithm, $\ln (Q^{2}/\mu _{c}^{2})$, is scheme dependent. \section{The Instanton-Induced Process\\ $\gamma^\ast + {\rm g}\rightarrow \overline{{\rm q}_L} + {\rm q}_R$} In this section, we turn to the central issue of this paper. We consider the simplest instanton-induced exclusive process, $\gamma^\ast + {\rm g}\rightarrow\overline{{\rm q}_L} + {\rm q}_R$, and compute its contributions to the fixed angle differential cross-section and the gluon structure functions ${\mathcal F}_{2\,{ g}}$ and ${\mathcal F}_{L\,{ g}}$, in leading semi-classical approximation. To this end, the respective Green's function is first set up according to standard instanton-perturbation theory in Euclidean configuration space~\cite{th,brown,ber,abc,r}, then Fourier transformed to momentum space, LSZ amputated, and finally continued to Minkowski space. The basic building blocks are (in Euclidean configuration space and in the singular gauge): \bigskip \noindent i) The classical instanton gauge field~\cite{bpst} $A^{(I)}_{\mu ^{\prime}}$, \begin{eqnarray} A_{\mu ^{\prime}}^{(I)}(x)&=&-\ii\,\frac{2\,\pi^{2}}{g_{s}}\, \rho^{2}\,U\,\left( \frac{ \sigma_{\mu ^{\prime}}\,\overline{x}-x_{\mu ^{\prime}}}{2\,\pi^{2}\,x^{4}} \right)\,U^{\dagger} \ \frac{1}{\Pi_{x}}\, , \label{igauge} \\ \Pi_x &\equiv & 1+\frac{\rho^2}{x^2} \, , \end{eqnarray} depending on the various collective coordinates, the instanton size $\rho$ and the colour orientation matrices $U^k_{\ \alpha}$. The $U$ matrices involve both colour ($k=1,2$) and spinor ($\alpha=1,2$) indices, the former ranging as usual only in the $2\times 2$ upper left corner of $3\times 3$ SU(3) colour matrices. Indices will, however, be suppressed, as long as no confusion can arise. \bigskip \noindent ii) The quark zero modes~\cite{th}, $\kappa$ and $\overline{\phi}$, \begin{eqnarray} \kappa^m_{\ \dot\beta}\,(x) &=& 2\,\pi\,\rho^{3/2}\, \epsilon^{\gamma\delta}\, \left( U\right)^m_{\ \delta}\, \frac{\overline{x}_{\dot\beta\gamma}}{2\,\pi^2\,x^4}\ \frac{1}{\Pi_x^{3/2}}\, , \label{kappa} \\ \overline{\phi}^{\dot\alpha}_{\ l}\,(x) &=&2\,\pi\,\rho^{3/2}\, \epsilon_{\gamma\delta}\, \left( U^\dagger\right)^\gamma_{\ l}\, \frac{x^{\delta\dot\alpha}}{2\,\pi^2\,x^4}\ \frac{1}{\Pi_x^{3/2}}\, , \label{phibar} \end{eqnarray} and \bigskip \noindent iii) the quark propagators in the instanton background~\cite{brown}, \begin{eqnarray} \label{si} &&S^{(I)}(x,y) = \\[1.6ex] \nonumber && \frac{1}{\sqrt{\Pi_x\Pi_y}} \left[ \frac{x-y}{2\pi^2(x-y)^4} \left( 1 +\rho^2\frac{ \left[ U x \overline{y} U^\dagger\right]} {x^2 y^2}\right) +\frac{\rho^2\sigma_\mu }{4\pi^2} \frac{\left[ U x\,(\overline{x}-\overline{y}) \sigma_\mu \overline{y} U^\dagger \right]} {x^2(x-y)^2 y^4\Pi_y} \right], \\[2ex] \label{sibar} &&\overline{S}^{(I)}(x,y) = \\[1.6ex] \nonumber && \frac{1}{\sqrt{\Pi_x\Pi_y}} \left[ \frac{\overline{x}-\overline{y}}{2\pi^2(x-y)^4} \left( 1 +\rho^2\frac{ \left[ U x \overline{y} U^\dagger\right]} {x^2 y^2}\right) +\frac{\rho^2 \overline{\sigma}_\mu }{4\pi^2} \frac{\left[ U x \overline{\sigma}_\mu (x-y) \overline{y} U^\dagger \right]} {\Pi_x x^4(x-y)^2 y^2 } \right] . \end{eqnarray} \begin{figure} \begin{center} \epsfig{file=ampl_ld.ps,width=9cm} \caption[dum]{\label{f3} Instanton-induced {\it chirality-violating} process, $\gamma^{\ast}(q) +{\rm g}(p )\rightarrow \overline{{\rm q}_{L}}(k_{1})+{\rm q}_{R}(k_{2})$, in leading semi-classical approximation. The corresponding Green's function involves the products of the appropriate classical fields (lines ending at blobs) as well as the quark propagator in the instanton background (quark line with central blob). } \end{center} \end{figure} The relevant diagrams for the exclusive process of interest, Eq.~(\ref{instproc}), are displayed in Fig.~\ref{f3}, in leading semi-classical approximation. The amplitude is expressed in terms of an integral over the collective coordinates $\rho$ and the colour orientation $U$, \begin{equation} {\mathcal T}_{\mu \,\mu ^{\prime}}^{a} \left( \gamma^\ast + {\rm g}\rightarrow \overline{{\rm q}_L} + {\rm q}_R \right) = \int\limits_0^\infty \frac{d\rho}{\rho^5}\,d(\rho ,\mu _{r} )\,\int dU\, {\mathcal A}_{\mu \,\mu ^{\prime}}^{a}(\rho, U) \,;\ a=1,2,3, \label{ampl_inst} \end{equation} where \begin{equation} d(\rho ,\mu _{r} )= d\,\left( \frac{2\,\pi}{\alpha_s(\mu _{r} )}\right)^6\, {\exp}\left[-\frac{2\,\pi}{\alpha_s(\mu _{r} )}\right] \, \left( \rho\,\mu _{r} \right)^{\beta_0+\frac{\alpha_s(\mu _{r} )}{4\,\pi}\, \left( \beta_1 - 12\,\beta_0 \right)}\, , \label{density} \end{equation} denotes the instanton density~\cite{th,ber,abc,morretal}, with $\mu _{r}$ being the renormalization scale. The form (\ref{density}) of the density, with next-to-leading order expression for $\alpha_{s}(\mu _{r})$, \begin{equation} \label{alpha} \alpha_{s}(\mu _{r})= \frac{4\,\pi}{\beta_{0}\,\ln \left(\frac{\mu _{r}^{2}}{\Lambda^{2}}\right)} \left[ 1 - \frac{\beta_{1}}{\beta_{0}^{2}} \frac{\ln\left(\ln \left(\frac{\mu _{r}^{2}}{\Lambda^{2}}\right) \right)} {\ln \left(\frac{\mu _{r}^{2}}{\Lambda^{2}}\right)} \right] \, , \end{equation} is improved to satisfy renormalization-group invariance at the 2-loop le\-vel~\cite{morretal}. The constants $\beta_{0}$ and $\beta_{1}$ are the familiar perturbative coefficients of the QCD beta-function, \begin{equation} \label{beta} \beta_0=11-\frac{2}{3}\,n_f\, ;\hspace{24pt} \beta_1=102-\frac{38}{3}\,n_f \, , \end{equation} and the constant $d$ is given by\footnote{Strictly speaking, the constant $d$ is known only to $1$-loop accuracy. In Ref.~\cite{morretal}, only the ultraviolet divergent part of the $2$-loop correction to the instanton density has been computed.} \begin{equation} \label{d} d=\frac{C_1}{2}\,{\rm e}^{-3\,C_2+n_f\,C_3}\, , \end{equation} with $C_{1}=0.466$, $C_{2}=1.54$, and $C_{3}=0.153$, in the $\overline{\rm MS}$-scheme. In our case, we should of course take $n_{f}=1$ in Eqs.~(\ref{beta}) and (\ref{d}). Before analytic continuation, the amplitude ${\mathcal A}_{\mu \,\mu ^{\prime}}$ entering Eq.~(\ref{ampl_inst}) takes the following form in Euclidean space, \vfill\eject \begin{eqnarray} \label{ampi} \lefteqn{ {\mathcal A}^{a}_{\mu \,\mu ^{\prime}} = -\ii\,e_{q}\, \lim_{p^{2}\to 0}p^{2}\, {\rm tr}\left( \sigma^{a}\,A_{\mu ^{\prime}}^{(I)}(p)\right)\times } \\ \nonumber && \chi_{R}^{\dagger}(k_{2})\, \left[ \lim_{k_{2}^{2}\to 0}\,(\ii k_{2})\,\kappa (-k_{2})\, {{\mathcal V}_\mu ^{(t)}}(q,-k_{1})\right . \\ \mbox{}&&+ \left . {{\mathcal V}_\mu ^{(u)}}(q,-k_{2})\,\lim_{k_{1}^{2}\to 0}\, \overline{\phi}(-k_{1})\,(-\ii\,\overline{k}_{1}) \right] \chi_{L}(k_{1})\, , \nonumber \end{eqnarray} with contributions ${{\mathcal V}_\mu ^{(t,u)}}$ from the diagrams on the left and right in Fig.~\ref{f3}, respectively, \begin{eqnarray} {{\mathcal V}_\mu ^{(t)}}(q,-k_{1})&\equiv& \int d^{4}x\,{\rm e}^{-\ii\,q\cdot x} \, \left[ \overline{\phi}(x)\,\overline{\sigma}_\mu \, \lim_{k_{1}^{2}\to 0}\,S^{(I)}\,(x,-k_{1})\,(-\ii\,\overline{k}_{1})\, \right] , \label{tinst}\\ {{\mathcal V}_\mu ^{(u)}}(q,-k_{2})&\equiv& \int d^{4}x\,{\rm e}^{-\ii\,q\cdot x} \, \left[ \lim_{k_{2}^{2}\to 0}\,(\ii k_{2})\,\overline{S}^{(I)}\,(-k_{2},x)\, \sigma_\mu \,\kappa(x)\right]\, , \label{uinst} \end{eqnarray} and generic notation for various Fourier transforms involved, \begin{equation} \label{fourier} f(\ldots,k,\ldots)=\int d^4 x\,{\rm e}^{-\ii\,k\cdot x}\, f(\ldots,x,\ldots)\, . \end{equation} The LSZ-amputation of the classical instanton gauge field $A^{(I)}_{\mu ^{\prime}}$ in Eq.~(\ref{ampi}) and the quark zero modes $\kappa$ and $\overline{\phi}$ in Eqs.~(\ref{tinst}) and (\ref{uinst}), respectively, is straightforward~\cite{r}, \begin{eqnarray} \lim_{p^{2}\to 0}p^{2}\, {\rm tr}\left( \sigma^{a}\,A_{\mu ^{\prime}}^{(I)}(p)\right) &=& \frac{2\,\pi^{2}}{g_{s}}\,\rho^{2}\, {\rm tr}\left[ \sigma^{a}\,U\,\left[ p_{\mu ^{\prime}}-\sigma_{\mu ^{\prime}}\,\overline{p}\right]\,U^{\dagger} \right]\, , \label{lszgluon}\\ \lim_{k_{2}^{2}\to 0}\, (\ii k_{2})^{\alpha\dot\alpha}\,\kappa^{i}_{\dot\alpha}(-k_{2})&=& 2\,\pi\,\rho^{3/2}\,U^{i}_{\ \beta}\,\epsilon^{\beta\alpha}\, , \label{lszkappa}\\ \lim_{k_{1}^{2}\to 0}\,\overline{\phi}_{j}^{\dot\gamma}(-k_{1})\, (-\ii\,\overline{k}_{1})_{\dot\gamma\delta} &=&2\,\pi\,\rho^{3/2}\,\epsilon_{\beta\delta}\, \left( U^{\dagger}\right)^{\beta}_{\ j}\, . \label{lszbarphi} \end{eqnarray} On the other hand, the LSZ-amputation of the quark propagators $S^{(I)}$ and $\overline{S}^{(I)}$ in Eqs.~(\ref{tinst}) and (\ref{uinst}), respectively, is quite non-trivial and has important physical consequences, as we shall see below. We give here only the final result and refer the interested reader to Appendix A where the details of the calculation can be found: \vfill\eject \begin{eqnarray} \label{onshellsi} \lefteqn{ \lim_{k_1^2\to 0}\, S^{(I)}\,(x,-k_1)\,(-\ii\, \overline{k_1}) =} \\[0.8ex] \nonumber && \frac{- 1}{\sqrt{\Pi_x}}\, {\rm e}^{\ii\,k_1\cdot x}\, \left[ 1 + \frac{1}{2}\,\frac{\rho^2}{x^2}\, \frac{\left[ U\,x\,\overline{k_1}\,U^\dagger \right]}{k_1\cdot x}\, \left( 1- {\rm e}^{-\ii\,k_1\cdot x}\right) \right] \, , \\[1.6ex] \label{onshellsibar} \lefteqn{ \lim_{k_2^2\to 0}\, \left( \ii\,k_2\right)\, \overline{S}^{(I)}\,(-k_2,x) =} \\[0.8ex] \nonumber && \frac{- 1}{\sqrt{\Pi_x}}\, {\rm e}^{\ii\,k_2\cdot x}\, \left[ 1 + \frac{1}{2}\,\frac{\rho^2}{x^2}\, \frac{\left[ U\,k_2\,\overline{x}\,U^\dagger \right]}{k_2\cdot x}\, \left( 1- {\rm e}^{-\ii\,k_2\cdot x}\right) \right] \, . \end{eqnarray} It should be noted that the first terms in Eqs.~(\ref{onshellsi}), (\ref{onshellsibar}), corresponding to the $1$ in square brackets, were argued to be present on general grounds already in Ref.~\cite{bbb}. The remaining terms, however, have not been given in the literature. As we shall see below, they play a very important r{\^o}le in ensuring electromagnetic gauge invariance. The Fourier transforms entering Eqs.~(\ref{tinst}) and (\ref{uinst}), respectively, can now be done with the help of Eqs.~(\ref{onshellsi}) and (\ref{onshellsibar}). The result is (see Appendix B), \begin{eqnarray} \label{tvertex} \lefteqn{ {\mathcal V}^{(t)}_{\mu \ j\,\alpha}\,(q,-k_{1}) = 2\,\pi\,\ii\,\rho^{3/2}\, \left( U^{\dagger}\right)^\gamma_{\ j}\, \left\{ \frac{1}{2}\, \frac{\left[ \epsilon\,k_{1}\,\overline{\sigma}_\mu \right]_{\gamma\alpha}} {q\cdot k_{1}}\, f\left( \rho\,\sqrt{q^{2}} \right) \right .} \\[1.6ex] \nonumber && \left . + \left[ \frac{\left[ \epsilon\,\left(q-k_{1}\right) \,\overline{\sigma}_\mu \right]_{\gamma\alpha}} {(q-k_{1})^2} -\frac{1}{2}\, \frac{\left[ \epsilon\,k_{1}\,\overline{\sigma}_\mu \right]_{\gamma\alpha}} {q\cdot k_{1}} \right]\,f\left(\rho\,\sqrt{\left( q-k_{1}\right)^2}\right) \right\}\, , \nonumber \\[2.4ex] \label{uvertex} \lefteqn{ {\mathcal V}^{(u)\ \alpha\, i}_{\mu \ }\,(q,-k_{2}) = 2\,\pi\,\ii\,\rho^{3/2}\, U^i_{\ \gamma}\, \left\{ \frac{1}{2}\, \frac{\left[ \sigma_\mu \, \overline{k_{2}}\,\epsilon\right]^{\alpha\gamma}} {q\cdot k_{2}}\, f\left( \rho\,\sqrt{q^{2}}\right) \right .} \\[1.6ex] \nonumber && \left . + \left[ \frac{\left[ \sigma_\mu \, \left(\overline{q}-\overline{k_{2}}\right)\,\epsilon \right]^{\alpha\gamma}} {(q-k_{2})^2} -\frac{1}{2}\, \frac{\left[ \sigma_\mu \, \overline{k_{2}}\,\epsilon\right]^{\alpha\gamma}} {q\cdot k_{2}} \right]\,f\left(\rho\,\sqrt{\left( q-k_{2}\right)^2}\right) \right\}\, , \nonumber \end{eqnarray} with the shorthand (``form factor''), \begin{equation} \label{form} f(\omega )\equiv \omega\,K_{1}(\omega ), \end{equation} in terms of the Bessel-K function, implying the normalization, \begin{equation} \label{formnorm} f(0)=1\, . \end{equation} The next step is to insert Eqs.~(\ref{tvertex}), (\ref{uvertex}), and (\ref{lszgluon}-\ref{lszbarphi}) into Eq.~(\ref{ampi}) and to perform the integration over the colour orientation according to Eq.~(\ref{ampl_inst}) by means of the relation, \begin{eqnarray} \label{colorint} \lefteqn{ \int dU\,U^k_{\ \beta^\prime}\,(U^\dagger )^{\gamma^\prime}_{\ l}\, U^i_{\ \tau}\,(U^\dagger )^\gamma_{\ j} =} \\ \nonumber &&\frac{1}{6}\,\left[ \delta_\tau^{\ \gamma^\prime}\,\delta^i_{\ l}\ \delta_{\beta^\prime}^{\ \gamma}\,\delta^k_{\ j} + \delta_\tau^{\ \gamma}\,\delta^i_{\ j}\ \delta_{\beta^\prime}^{\ \gamma^\prime}\,\delta^k_{\ l} + \epsilon_{\tau\,\beta^\prime}\,\epsilon^{i\,k}\ \epsilon^{\gamma^\prime\,\gamma}\,\epsilon_{l\,j} \right] \, . \end{eqnarray} After analytic continuation to Minkowski space we find for the scattering amplitude, Eq.~(\ref{ampl_inst}), \begin{eqnarray} \label{ampi2} && {\mathcal T}_{\mu \,\mu ^{\prime}}^a\, \left( \gamma^\ast + {\rm g}\rightarrow \overline{{\rm q}_L} + {\rm q}_R \right) = -\ii\,\frac{4}{3}\,\pi^4\,\frac{e_q}{g_s}\,\sigma^a \int\limits_0^\infty d\rho\,d(\rho ,\mu _{r} )\,\times \\[0.8ex] && \chi_R^\dagger (k_2) \left[ \left( \sigma_{\mu ^{\prime}}\overline{p}-p\overline{\sigma}_{\mu ^\prime}\right) V(q,k_1;\rho )\overline{\sigma}_\mu -\sigma_\mu \overline{V}(q,k_2;\rho ) \left( \sigma_{\mu ^\prime}\overline{p} - p\overline{\sigma}_{\mu ^\prime} \right) \right] \chi_L(k_1) , \nonumber \end{eqnarray} with the four-vector $V_{\lambda}$, \begin{eqnarray} \nonumber V_\lambda (q,k;\rho ) &\equiv& \left[ \frac{\left( q-k\right)_\lambda}{-(q-k)^2} +\frac{k_{\lambda}}{2 q\cdot k} \right]\rho\sqrt{-\left( q-k\right)^2}\, K_1\left(\rho\sqrt{-\left( q-k\right)^2}\right) \\[0.8ex] \label{V} \mbox{}&& - \frac{k_{\lambda}}{2 q\cdot k}\rho\sqrt{-q^{2}}\, K_1\left(\rho\sqrt{-q^{2}}\right) . \end{eqnarray} At this stage of our instanton calculation, the gauge-invariance con\-st\-raints, Eqs.~(\ref{gaugeinv}), can easily be checked. While the relation ${\mathcal T}_{\mu \,\mu ^{\prime}}^a\,p^{\mu ^\prime}=0$ holds trivially, the electromagnetic (e.m.) current conservation $q^\mu \,{\mathcal T}_{\mu \,\mu ^{\prime}}^a=0$ follows again from the relations (\ref{com}) of the $\sigma$-matrices, the Weyl-equations (\ref{weyleq}) and the on-shell conditions $k_1^2=k_2^2=0$. Electromagnetic current conservation provides also for a non-trivial check of our result for the amputated quark propagators, which differs somewhat from the result quoted in Ref.~\cite{bbb}: If we keep only the first terms in Eqs.~(\ref{onshellsi}), (\ref{onshellsibar}), corresponding to the $1$ in square brackets, e.m. current conservation would only hold for a restricted set of momenta in phase space, namely for $(q-k_1)^2=(q-k_2)^2$. Furthermore, we note one of the main results of this paper: The integration over the instanton size $\rho$ in Eq.~(\ref{ampi2}) is {\it finite}. In particular, the {\it good infrared} behavior (large $\rho$) of the integrand is due to the exponential decrease of the Bessel-K function for large $\rho$ in Eq.~(\ref{V}). Its origin, in turn, can be traced back to the ``feed-through'' of the factor $1/\sqrt{\Pi_x}$, by which the amputated (current) quark propagators (\ref{onshellsi}) and (\ref{onshellsibar}) in the $I$-background essentially differ from the respective amputated free propagators. If the current-quark propagators in Eqs.~(\ref{tinst}) and (\ref{uinst}) are naively approximated by the free ones (c.\,f. Eqs.~(\ref{si}), (\ref{sibar})), \begin{equation} S^{(0)}(x,y)=\frac{x-y}{2\pi^2(x-y)^4}\,;\hspace{24pt} \overline{S}^{(0)}(x,y)=\frac{\overline{x}-\overline{y}}{2\pi^2(x-y)^4}\, , \end{equation} the result is both gauge variant and contains an IR-divergent piece in the $\rho$ integration. We have thus demonstrated explicitly and to our knowledge for the first time that the typical hard scales ($Q^{2},\ldots$) in deep-inelastic scattering provide a dynamical IR cutoff for the instanton size (at least in leading semi-classical approximation). Now we are ready to perform the final integration over the instanton size $\rho$ by inserting the instanton density, Eq.~(\ref{density}), into Eq.~(\ref{ampi2}). The result is: \begin{eqnarray} \label{ampifin} && {\mathcal T}_{\mu \,\mu ^{\prime}}^a\, \left( \gamma^\ast + {\rm g}\rightarrow \overline{{\rm q}_L} + {\rm q}_R \right) = \\[2.4ex] \nonumber && -\ii\frac{\sqrt{2}}{3}d\pi^{3}e_q \left( \frac{2\,\pi}{\alpha_s(\mu _{r} )}\right)^{13/2} {\exp}\left[-\frac{2\,\pi}{\alpha_s(\mu _{r} )}\right]\, 2^{\,b}\Gamma \left(\frac{b+1}{2}\right) \Gamma \left(\frac{b+3}{2}\right) \sigma^a\,\times \\[2.4ex] && \chi_R^\dagger (k_2) \left[ \left( \sigma_{\mu ^{\prime}}\overline{p}-p\overline{\sigma}_{\mu ^\prime}\right) v(q,k_1;\mu _{r} )\overline{\sigma}_\mu -\sigma_\mu \overline{v}(q,k_2;\mu _{r} ) \left( \sigma_{\mu ^\prime}\overline{p} - p\overline{\sigma}_{\mu ^\prime} \right) \right] \chi_L(k_1) , \nonumber \end{eqnarray} with the four-vector $v_{\lambda}$, \begin{eqnarray} \label{v} &&v_\lambda (q,k;\mu _{r} ) \equiv \\[1.6ex] \nonumber && \frac{1}{\mu _{r}}\,\left\{ \left[ \frac{\left( q-k\right)_\lambda}{-(q-k)^2} +\frac{k_{\lambda}}{2 q\cdot k} \right] \left( \frac{\mu _{r}}{\sqrt{-\left( q-k\right)^2}}\right)^{b+1} - \frac{k_{\lambda}}{2 q\cdot k} \left( \frac{\mu _{r}}{\sqrt{-q^{2}}}\right)^{b+1} \right\} . \end{eqnarray} In Eqs.~(\ref{ampifin}) and (\ref{v}), the variable $b$ is a shorthand for the effective power of $\rho\mu _{r}$ in the instanton density, Eq.~(\ref{density}), \begin{equation} \label{beff} b\equiv \beta_0+\frac{\alpha_s(\mu _{r} )}{4\,\pi}\, \left( \beta_1 - 12\,\beta_0 \right) \, . \end{equation} The next steps consist in contracting the amplitude, Eq.~(\ref{ampifin}), with the gluon polarization vector $\epsilon_{g}^{\mu ^{\prime}}(p)$ and taking the modulus squared of this amplitude according to Eq.~(\ref{wmunu_n}). After applying Eq.~(\ref{compl}), the remaining spinor traces can be evaluated, in principle, by repeated use of Eq.~(\ref{com}). For the actual calculation, however, we used FORM and, for an independent check, the HIP package for MAPLE. The final result for the relevant projections (c.f. Eqs.~(\ref{diffcross}), (\ref{projF2}), and (\ref{projFL})) of the contribution of the $I$-induced process (\ref{instproc}) to the differential gluon structure tensor is found to be \begin{eqnarray} \nonumber \lefteqn{ -g_{\mu \nu}\, \frac{d{\mathcal W}^{\mu \nu\,{(2)}}_{g}}{dt} \, \left( \gamma^\ast + {\rm g}\rightarrow \overline{{\rm q}_L} + {\rm q}_R \right) =} \\[1.6ex] \label{gmunu} &&\frac{e_q^2}{16}\, {\mathcal N}^{2}\, \left( \frac{2\,\pi}{\alpha_s(\mu _{r} )}\right)^{13}\, {\exp}\left[-\frac{4\,\pi}{\alpha_s(\mu _{r} )}\right]\, \left( \frac{\mu _{r}^{2}}{Q^{2}}\right)^{b}\,\times \\[1.6ex] \nonumber && \frac{1-x}{Q^{2}}\, \left[ \left(\frac{Q^{2}}{-t}\right)^{b+1} + \left(\frac{Q^{2}}{-u}\right)^{b+1} + 2\,t\,u\, \frac{ \left( \left(\frac{Q^{2}}{-t}\right)^{\frac{b+1}{2}} -1 \right) \left( \left(\frac{Q^{2}}{-u}\right)^{\frac{b+1}{2}} -1 \right) } {(t+Q^{2})(u+Q^{2})} \right] \, , \\[2.4ex] \nonumber \lefteqn{ \frac{p_\mu \,p_{\nu}}{p\cdot q} \, \frac{d{\mathcal W}^{\mu \nu\,{(2)}}_{g}}{dt} \, \left( \gamma^\ast + {\rm g}\rightarrow \overline{{\rm q}_L} + {\rm q}_R \right) =} \\[1.6ex] \label{pmupnu} &&\frac{e_q^2}{16} {\mathcal N}^{2}\, \left( \frac{2\,\pi}{\alpha_s(\mu _{r} )}\right)^{13}\, {\exp}\left[-\frac{4\,\pi}{\alpha_s(\mu _{r} )}\right] \, \left( \frac{\mu _{r}^{2}}{Q^{2}}\right)^{b}\,\times \\[1.6ex] \nonumber && \frac{(1-x)^{2}}{Q^{2}}\, t\,u\, \left[ \frac{\left(\frac{Q^{2}}{-t}\right)^{\frac{b+1}{2}} +\frac{u}{Q^{2}}\frac{x}{1-x}}{t+Q^{2}} -\frac{\left(\frac{Q^{2}}{-u}\right)^{\frac{b+1}{2}} +\frac{t}{Q^{2}}\frac{x}{1-x}}{u+Q^{2}} \right]^{2} \, . \end{eqnarray} Here we have introduced the shorthand \begin{equation} {\mathcal N} \equiv \sqrt{\frac{2}{3}}\,\pi^{2}\,d\ 2^{\,b}\, \Gamma \left(\frac{b+1}{2}\right)\, \Gamma \left(\frac{b+3}{2}\right)\, . \end{equation} In Eqs.~(\ref{gmunu}) and (\ref{pmupnu}), the $t\leftrightarrow u$ symmetry is manifest. \begin{figure} \begin{center} \epsfig{file=dsigdz.eps,width=9cm}\vfill \hspace{8pt} \epsfig{file=fixangle.eps,width=8cm} \caption[dum]{\label{f4} Differential cross-sections, $d\sigma /d\cos\theta\ [\mbox{nb}]$, of the $I$-induced {\it chirality-violating} process (\ref{instproc}) and the perturbative {\it chirality-conserving} process (\ref{pertproc}), both for fixed c.m. scattering angle, $\theta=90^{\,\circ}$, $\Lambda=0.234$ GeV, $e_{q}=2/3$, and $\mu _{r}=Q$. Top: For fixed $x=0.25$, as function of $Q$ [GeV]. Bottom: For fixed $Q=10$ GeV, as function of $x$.} \end{center} \end{figure} Upon inserting Eqs.~(\ref{gmunu}), (\ref{pmupnu}) into Eqs.~(\ref{diffcross}), (\ref{projF2}), and (\ref{projFL}), we see that the contribution of the $I$-induced process (\ref{instproc}) to the differential cross-section $d\sigma /dt$ and the differential gluon structure functions, $d{\mathcal F}^{(2)}_{2\,g}/dt, d{\mathcal F}^{(2)}_{L\,{g}}/dt$, is well-behaved as long as we avoid the (collinear) singularities for $t,u\rightarrow 0$. This is illustrated in Fig.~\ref{f4}, where we compare the differential cross-sections, $d\sigma /\cos\theta$, of both the $I$-induced process, Eq.~(\ref{instproc}), and the perturbative process, Eq.~(\ref{pertproc}), where \begin{equation} \label{kt} t=-\frac{Q^{2}}{2\,x}\, \left( 1- \cos \theta \right) \, . \end{equation} We note that the renormalization-scale dependence of the $I$-induced cross-section in Fig.~\ref{f4} is very small, due to the re\-nor\-ma\-li\-za\-tion-group im\-proved density (\ref{density}). Let us address at this point the important question concerning the range of validity of the present calculation. Specifically, let us examine the constraints emerging from the requirement of the dilute instanton gas approximation following Refs.~\cite{cdg,ag,as,svz}. Along these lines one finds that instantons with size \begin{equation} \label{rhoc} \rho > \rho_{c}\simeq 1/(500\ {\rm MeV}) \end{equation} are ill-defined semi-classically~\cite{svz}, corresponding to a breakdown of the dilute gas approximation. On the other hand, using the form of our $\rho$ integral in Eqs.~(\ref{ampi2}), (\ref{V}), we may determine the average instanton size $\langle \rho\rangle$ contributing for a given virtuality \begin{equation} \label{virt} {\mathcal Q}=\min \left( Q, \sqrt{-t}=\frac{Q}{\sqrt{x}}\sin\frac{\theta}{2}, \sqrt{-u}=\frac{Q}{\sqrt{x}}\cos\frac{\theta}{2} \right)\, , \end{equation} according to \begin{equation} \label{avrho} \langle \rho\rangle \equiv \frac{\int\limits_{0}^{\infty}d\rho\, \rho\,\rho^{b+1}\,K_{1}(\rho\,{\mathcal Q})} {\int\limits_{0}^{\infty}d\rho\, \rho^{b+1}\,K_{1}(\rho\,{\mathcal Q})} \simeq \frac{b+3/2}{\mathcal Q} \, . \end{equation} Hence, with Eq.~(\ref{rhoc}) and Eq.~(\ref{beff}), we find that the virtuality $\mathcal Q$ should obey \begin{equation} \label{qmin} {\mathcal Q}\ (>)> (5-6)\ {\rm GeV}\, . \end{equation} In particular, our results in Fig.~\ref{f4} (top) for the $I$-induced differential cross-section, $d\sigma /d\cos\theta$, at $\theta=90^{\,\circ}$ and $x=0.25$, should be taken seriously only for $Q\ (= {\mathcal Q})> (5-6)$ GeV (since here $\sqrt{-t}=\sqrt{-u}=Q/\sqrt{2x}>Q$). Thus, like in the perturbative case, {\it fixed-angle scattering processes at high $Q^2$ are reliably calculable in (instanton) perturbation theory} (at least in leading semi-classical approximation). Next, we note that the contributions (\ref{gmunu}), (\ref{pmupnu}) of the $I$-induced exclusive process (\ref{instproc}) to the differential gluon structure functions are much more singular ($\propto (-t, -u)^{-(b+1)}$) for $t\to 0,u\to 0$ than the perturbative ones ($\propto (-t, -u)^{-1} $). This leads to a much stronger scheme dependence~\cite{mrs} than in the perturbative case. Let us have a closer look at this feature. We regularize the collinear divergence of the $t$ integral along the same lines as in perturbation theory, i.e. we restrict the integration to the interval $\{ -Q^{2}/x+\mu _{c}^{2},-\mu _{c}^{2}\}$. On account of Eqs.~(\ref{projF2}), (\ref{projFL}), we then obtain for the hard contributions of the $I$-induced exclusive process (\ref{instproc}) to the gluon structure functions, \begin{eqnarray} \nonumber {\mathcal F}_{2\,g}^{(\overline{{\rm q}_L}{\rm q}_R)}\,(x,Q^2;\mu _{c}^{2}) & =&\frac{e_q^2}{8}\, {\mathcal N}^{2}\, \left( \frac{2\,\pi}{\alpha_s(\mu _{r} )}\right)^{13}\, {\exp}\left[-\frac{4\,\pi}{\alpha_s(\mu _{r} )}\right] \,\left( \frac{\mu _{r}^{2}}{\mu _{c}^{2}}\right)^{b} \\[1.6ex] \label{F2inst} \mbox{}&& \times \frac{x\,(1-x)}{b}\, \left[1+{\mathcal O}\left( \frac{\mu _{c}^{2}}{Q^{2}}\right)\right] \, , \\[2.4ex] \nonumber {\mathcal F}_{L\,g}^{(\overline{{\rm q}_L}{\rm q}_R)}\,(x,Q^2;\mu _{c}^{2}) & =&\frac{e_q^2}{2}\, {\mathcal N}^{2}\, \left( \frac{2\,\pi}{\alpha_s(\mu _{r} )}\right)^{13}\, {\exp}\left[-\frac{4\,\pi}{\alpha_s(\mu _{r} )}\right] \,\left( \frac{\mu _{r}^{2}}{\mu _{c}^{2}}\right)^{b}\, \frac{\mu _{c}^{2}}{Q^{2}}\, \\[1.6ex] \label{FLinst} \mbox{}&& \times \, \frac{x\,(1-x)^{2}}{b-1}\, \left[1+{\mathcal O}\left(\frac{\mu _{c}^{2}}{Q^{2}}\right)\right] . \end{eqnarray} In the Bjorken limit, $Q^{2}/\mu _{c}^{2}\to \infty$, but with $\mu _{c}^{2}/\mu _{r}^{2}$ fixed, we find from Eqs.~(\ref{F2inst}) and (\ref{FLinst}), respectively, \begin{eqnarray} \label{F2instbj} \lefteqn{ \lim_{Q^{2}\to \infty}\, {\mathcal F}_{2\,g}^{(\overline{{\rm q}_L}{\rm q}_R)}(x,Q^2;\mu _{c}^{2}) =} \\[1.6ex] \nonumber && \frac{e_q^2}{8}\, {\mathcal N}^{2}\, \left( \frac{2\,\pi}{\alpha_s(\mu _{r} )}\right)^{13}\, {\exp}\left[-\frac{4\,\pi}{\alpha_s(\mu _{r} )}\right] \,\left( \frac{\mu _{r}^{2}}{\mu _{c}^{2}}\right)^{b}\, \frac{x\,(1-x)}{b} \, , \\[2.4ex] \label{FLinstbj} \lefteqn{ \lim_{Q^{2}\to \infty}\, {\mathcal F}_{L\,g}^{(\overline{{\rm q}_L}{\rm q}_R)}(x,Q^2;\mu _{c}^{2}) \equiv } \\[1.6ex] \nonumber &&\lim_{Q^{2}\to \infty}\,\left[ {\mathcal F}_{2\,g}^{(\overline{{\rm q}_L}{\rm q}_R)} (x,Q^2;\mu _{c}^{2}) -2\,x\,{\mathcal F}_{1\,g}^{(\overline{{\rm q}_L}{\rm q}_R)} (x,Q^2;\mu _{c}^{2})\right] = 0\, . \end{eqnarray} Hence, in this limit, the considered $I$-induced process gives a {\it scaling} contribution to ${\mathcal F}_{2\,g}$ and the analogue of the Callan-Gross relation, ${\mathcal F}_{2\,g}=2\,x\,{\mathcal F}_{1\,g}$, holds. In particular, this means that the {\it same} parton distribution can absorb the infrared sensitivity of both structure functions, ${\mathcal F}_{2\,g}^{(\overline{{\rm q}_L}{\rm q}_R)}$ and ${\mathcal F}_{1\,g}^{(\overline{{\rm q}_L}{\rm q}_R)}$. This is one of the prerequisites of factorization~\cite{pert}. \section{Conclusions and Outlook} In this paper, we studied QCD-instanton induced processes in deep-inelastic lepton-hadron scattering. The purpose of the present work was to shed further light on the important questions around the IR-behaviour associated with the instanton size. In order to eliminate unessential technical complications, we have reduced the realistic task of evaluating the $I$-induced cross-sections of the chirality violating multi-particle processes (illustrated in Fig.~\ref{f1}) \begin{equation} \gamma^{\ast}+{\rm g}\Rightarrow \overline{{\rm u}_{L}}+{\rm u}_{R}\, + \overline{{\rm d}_{L}}+{\rm d}_{R}\, +\overline{{\rm s}_{L}}+{\rm s}_{R}\, +n_g\, {\rm g}, \end{equation} to the detailed study of the {\it simplest} one, without additional gluons and with just one massless flavour ($n_f=1$), \begin{equation} \gamma^{\ast}(q)+{\rm g}(p)\Rightarrow \overline{{\rm q}_{L}}(k_{1})+{\rm q}_{R}(k_{2})\, . \end{equation} We have explicitly calculated the corresponding fixed angle cross-section and the contributions to the gluon structure functions within standard instanton perturbation theory in leading semi-classical approximation (Sect.~3). To this approximation, {\it fixed-angle scattering processes at high $Q^2$ are reliably calculable}. In the Bjorken limit, the considered $I$-induced process gives a {\it scaling} contribution to ${\mathcal F}_{2\,g}$ and the analogue of the Callan-Gross relation, ${\mathcal F}_{2\,g}=2\,x\,{\mathcal F}_{1\,g}$, holds. All along we focused our main attention on the IR behaviour associated with the instanton size. Gauge invariance was kept manifest along the calculation. As a central result of this paper and unlike Ref.~\cite{bb}, we found {\it no} IR divergencies associated with the integration over the instanton size $\rho$, which can even be perfomed analytically. We have explicitly demonstrated that the typical hard scale ${\mathcal Q}$ in deep-inelastic scattering provides a {\it dynamical} infrared cutoff for the instanton size, $\rho \mbox{\,\raisebox{.3ex {\mathcal O}(1/{\mathcal Q})$. Thus, deep-inelastic scattering may indeed be viewed as a distinguished process for studying manifestations of QCD-instantons. In Ref.~\cite{bb}, the $I$-induced contribution to deep-inelastic scattering of a virtual gluon from a real one~\cite{bbgg}, $g^\ast g \Rightarrow g^\ast g$, served as a simplified r\^{o}le model for the splitting into a IR-finite contribution ($\rho \mbox{\,\raisebox{.3ex 1/Q$) and an IR-divergent term (large $\rho$). As speculated by one of the authors~\cite{b}, the occurence of the IR-divergent term could well have been due to the lacking gauge invariance of this model, associated with the off-shellness of one of the initial gluons. In fact, one may enforce gauge invariance by replacing the instanton gauge field $A^{(I)}_\mu (x)$ describing the virtual gluon by the familiar gauge-invariant operator \begin{equation} G^{(I)}_\mu (x)=e^\alpha\, [x+\infty,x]\, G^{(I)}_{\mu \,\alpha}(x)\,[x,x+\infty], \end{equation} where $G^{(I)}_{\mu \,\alpha}(x)$ is the instanton field-strength, and \begin{equation} [x,x+\infty]=P \exp\left\{ \ii\,g_{s} \int^\infty_0 d\lambda\,\, e\cdot A^{(I)}(x+\lambda e) \right\}\, , \end{equation} is a gauge factor ordered along the lightlike line in the direction $e_\mu =(q_\mu +x p_\mu )/(2p\cdot q)$. In this case the IR-divergent term is, indeed, absent~\cite{b}. Since our present calculation is manifestly gauge-invariant, the absence of an IR divergent term fits well in line with these arguments. A further main purpose of the present calculation was to provide a clean and explicit discussion of most of the crucial steps involved in our subsequent task~\cite{mrs} to calculate the {\it dominant} $I$-induced contributions coming from final states with a large number of gluons (and three massless flavours, say). Let us close with some comments on the generalization to the more realistic case with $n_g$ gluons in the final state, which is entirely straightforward~\cite{mrs,q96}. Instead of Eq.~(\ref{ampi2}), the corresponding amplitude involves, in leading semi-classical approximation, the additional factors from the $n_g$ gluons (c.f. Eq.~(\ref{lszgluon})), \begin{eqnarray} \label{ampi2g} \lefteqn{ {\mathcal T}_\mu ^{a\,a_1\ldots a_{n_g}}\,\left( \gamma^\ast + {\rm g}\rightarrow \overline{{\rm q}_L} + {\rm q}_R +n_g\,{\rm g}\right) = } \\[1.6ex] \nonumber && \ii\,e_q\,4\,\pi^2\,\left( \frac{\pi^3}{\alpha_s}\right)^{\frac{n_g+1}{2}}\, \int dU \int\limits_0^\infty d\rho\,d(\rho ,\mu _{r} )\,\rho^{2\,n_g} \\[1.6ex] \nonumber &&\times \, {\rm tr}\left[ \sigma^{a}\,U\,\left[ \epsilon_g(p)\cdot p-\epsilon_g(p)\,\overline{p} \right]\,U^{\dagger} \right]\, \prod_{i=1}^{n_g} {\rm tr}\left[ \sigma^{a_i}\,U\,\left[ \epsilon_g(p_i)\,\overline{p_i}-\epsilon_g(p_i)\cdot p_i \right]\,U^{\dagger} \right] \\[1.6ex] \nonumber &&\times \left\{ \left[ U \chi_R^\dagger (k_2 ) \epsilon \right] \left[\epsilon V(q,k_1;\rho)\overline{\sigma}_\mu \chi_L(k_1)\, U^\dagger \right] \right . \\[1.6ex] \nonumber \mbox{}&& - \left . \left[ U \chi_R^\dagger (k_2) \sigma_\mu \overline{V}(q,k_2;\rho)\epsilon \right] \left[ \epsilon\, \chi_L(k_1)\, U^\dagger \right] \right\} , \end{eqnarray} where the four-vector $V_\lambda$ is again given by Eq.~(\ref{V}). Besides the enhancement by a factor of $(\pi^3/\alpha_s)^{1/2}$, each additional gluon gives rise to a factor of $\rho^2$ under the $I$-size integral. The IR-finiteness of this integral is, however, not altered by the presence of the additional overall factor of $\rho^{2\,n_g}$, on account of the exponential cutoff $\propto \exp[ -\rho {\mathcal Q}]$ from the Bessel-K function in the ``form-factors'' contained in $V_\lambda(q,k;\rho)$, Eq.~(\ref{V}). We also note, that the amplitude (\ref{ampi2g}) satisfies e.m. gauge invariance. In analogy to electro-weak $(B+L)$-violation~\cite{m}, one expects~\cite{bb,rs} the sum of the final-state gluon contributions to exponentiate, such that the total $I$-induced $\gamma^\ast$g cross-section takes the form (at large $Q^2$), \begin{eqnarray} \label{exp} \sigma^{{ (I)}}_{\gamma^\ast { g}}(x,Q^2) &\equiv& \sum_{n_{ g}}\sigma^{{ (I)}}_{\gamma^\ast { g}\,n_{ g}}(x,Q^2) \\[1.6ex] \nonumber &\sim& \int_x^1 dx^\prime\int\limits^{Q^2\frac{{x^\prime}}{x}} \frac{dQ^{\prime 2}}{Q^{\prime 2}}\,\ldots\, \frac{1}{Q^{\prime 2}}\, \exp\left[-\frac{4\pi}{\alpha_s(Q^\prime)}\,F(x^\prime )\right]\, , \end{eqnarray} where the so-called ``holy-grail function''~\cite{m} $F({x^\prime} )$ (normalized to F(1)=1) is expected to decrease towards smaller ${x^\prime}$, which implies a dramatic growth of $\sigma^{{ (I)}}_{\gamma^\ast { g}}(x,Q^2)$ for decreasing $x$. \section*{Appendix A} Here we want to derive Eqs.~(\ref{onshellsi}) and (\ref{onshellsibar}) for the LSZ-amputated quark propagators. Let us first consider the Fourier transform of the quark propagator (\ref{si}) which we write as \begin{equation} S^{(I)}(x,-k)=\frac{1}{\sqrt{\Pi_x}} \sum_{i=1}^3 s^{(i)} (x,-k) \, , \label{sumsi} \end{equation} where \begin{eqnarray} \label{s1} \lefteqn{ s^{(1)}(x,-k) = } \\[1.6ex] \nonumber && 2\left[ x- (-\ii\partial_k)\right]\, (-\ii\overline{\partial}_k)\,(-\ii\partial_k)\, \int \frac{d^4y}{(2\pi )^2} \frac{{\rm e}^{\ii k\cdot y}} {\left( (x-y)^2\right)^2 \left( y^2+\rho^2\right)^{1/2} \left( y^2\right)^{1/2}} \, , \\[2.4ex] \label{s2} \lefteqn{ s^{(2)}(x,-k) =} \\[1.6ex] \nonumber && 2\,\frac{\rho^2}{x^2}\left[ x- (-\ii\partial_k)\right] U x (-\ii\overline{\partial}_k)\, U^\dagger \int \frac{d^4y}{(2\pi )^2} \frac{{\rm e}^{\ii k\cdot y}} {\left( (x-y)^2\right)^2 \left( y^2+\rho^2\right)^{1/2} \left( y^2\right)^{1/2}} \, , \\[2.4ex] \label{s3} \lefteqn{ s^{(3)}(x,-k) =} \\[1.6ex] \nonumber && \frac{\rho^2}{x^2} \sigma_\mu U x\, \left[ \overline{x}- (-\ii\overline{\partial}_k)\right] \sigma_\mu (-\ii\overline{\partial}_k)\, U^\dagger \int \frac{d^4y}{(2\pi )^2} \frac{{\rm e}^{\ii k\cdot y}} {\left( x-y\right)^2 \left( y^2+\rho^2\right)^{3/2} \left( y^2\right)^{1/2}} \, . \end{eqnarray} Our strategy to analyze the $k^2\to 0$ limit of Eqs.~(\ref{s1}-\ref{s3}) starts by partially evaluating the master integral, \begin{equation} I\,(-k;x,\rho ;\alpha , \beta , \gamma ) \equiv \int \frac{d^4y}{(2\pi )^2} \frac{{\rm e}^{\ii k\cdot y}} {\left( (x-y)^2\right)^\alpha \left( y^2+\rho^2\right)^\beta \left( y^2\right)^\gamma} \, , \label{masterint} \end{equation} by means of the Feynman parametrization (see e.g. \cite{yn}), \begin{eqnarray} \label{feynman} \lefteqn{ \frac{1}{A^\alpha B^\beta C^\gamma} =} \\[1.6ex] \nonumber && \frac{\Gamma (\alpha +\beta +\gamma ) } {\Gamma (\alpha )\,\Gamma (\beta )\, \Gamma (\gamma )}\, \int_0^1 da\,a \, \int_0^1 db\, \frac{(ab)^{\alpha -1} (a(1-b))^{\beta -1}\,(1-\alpha)^{\gamma -1}} { \left[ ab\,A +a\,(1-b)\,B + (1-a)\,C\right]^{\alpha +\beta +\gamma}} \, . \end{eqnarray} With the help of Eq.~(\ref{feynman}), it is possible to show that Eq.~(\ref{masterint}) can be expressed as \begin{eqnarray} \label{masterintfin} \lefteqn{ I\,(-k;x,\rho ;\alpha , \beta , \gamma ) =} \\[1.6ex] \nonumber && \frac{2^{1-(\alpha +\beta +\gamma )}}{ \Gamma (\alpha )\, \Gamma (\beta )\, \Gamma (\gamma )}\, \int_0^1 da\,a^{\alpha +\beta -1}\,(1-a)^{\gamma -1} \int_0^1 db\,b^{\alpha -1}\,(1-b)^{\beta -1}\, {\rm e}^{\ii\,k\cdot x\,ab} \\[1.6ex] \nonumber && \times \,\left( \frac{\sqrt{x^2 ab(1-ab)+\rho^2 a(1-b)}}{\sqrt{k^2}} \right)^{2-(\alpha +\beta +\gamma )} \\[1.6ex] \nonumber && \times \,K_{2-(\alpha +\beta +\gamma )} \left( \sqrt{k^2}\sqrt{x^2 ab(1-ab)+\rho^2\, a(1-b)} \right) . \end{eqnarray} Next we insert Eq.~(\ref{masterintfin}) into Eqs.~(\ref{s1})-(\ref{s3}), perform the various derivatives, and expand the integrand with respect to $k^2\to 0$. Finally, the remaining Feynman parameter integrations are done. After this procedure we find: \vfill\eject \begin{eqnarray} \lim_{k^{2}\to 0}{s^{(1)}} (x,-k)(-\ii\,\overline{k}) &=& -{\rm e}^{\ii\,k\cdot x}\, , \\[1.6ex] \lim_{k^{2}\to 0}{s^{(2)}} (x,-k)(-\ii\,\overline{k}) &=&- \frac{1}{2}\frac{\rho^2}{x^2} \frac{\left[ U x \overline{k} U^\dagger \right]}{k\cdot x} \left[ {\rm e}^{\ii\,k\cdot x} - \frac{\ii}{k\cdot x} \left(1-{\rm e}^{\ii\,k\cdot x}\right)\right] , \\[1.6ex] \lim_{k^{2}\to 0}{s^{(3)}} (x,-k)(-\ii\,\overline{k}) &=& \frac{1}{2}\frac{\rho^2}{x^2} \frac{\left[ U x\overline{k} U^\dagger \right]}{k\cdot x} \left[ 1 - \frac{\ii}{k\cdot x} \left(1-{\rm e}^{\ii\,k\cdot x}\right)\right] . \end{eqnarray} Thus, on account of Eq.~(\ref{sumsi}), the on-shell residuum of the quark propagator (\ref{si}) is given by Eq.~(\ref{onshellsi}). A similar reasoning leads to Eq.~(\ref{onshellsibar}) for the residuum of the quark propagator (\ref{sibar}). \section*{Appendix B} Our task is to derive Eqs.~(\ref{tvertex}), (\ref{uvertex}), corresponding to the $\gamma^{\ast}$- quark vertex ${\mathcal V}_\mu ^{(t,u)}$ in the leading-order $I$-induced amplitude. We will concentrate on the derivation of Eq.~(\ref{tvertex}), since the derivation of Eq.~(\ref{uvertex}) is completely analogous. Let us recall the definition of ${\mathcal V}_\mu ^{(t)}$, but now with indices written explicitly, \begin{eqnarray} {\mathcal V}^{(t)}_{\mu \ m \lambda}\,(q,-k) = \int d^4x\,{\rm e}^{-\ii\,q\cdot x}\ \overline{\phi}^{\dot\alpha}_l (x)\, \overline{\sigma}_{\mu \,\dot\alpha\alpha}\, \lim_{k^2\to 0}\, {{S}^{(I)\,\alpha\dot\beta\ l}}_m\,(x,-k)\, \left(-\ii\,\overline{k}_{\dot\beta\lambda}\right) . \label{phivertex1} \end{eqnarray} Inserting Eqs.~(\ref{phibar}) and (\ref{onshellsi}) into Eq.~(\ref{phivertex1}) we obtain for the vertex, \begin{eqnarray} {\mathcal V}^{(t)}_{\mu \ m \lambda}\,(q,-k) &=& -\frac{\rho^{3/2}}{\pi}\, \int d^4x\,{\rm e}^{-\ii\,( q-k)\cdot x}\, \frac{1}{\left( x^2+\rho^2\right)^2} \label{phivertex2} \\ &\times & \epsilon_{\gamma\delta}\, \left[ x\,\overline{\sigma}_\mu \right]^\delta_{\ \lambda}\, \left[ \left( U^\dagger\right)^\gamma_{\ m} +\frac{1}{2}\,\frac{\rho^2}{x^2}\, \frac{\left[ x\,\overline{k}\,U^\dagger\right]^\gamma_{\ m}} {k\cdot x}\, \left( 1-{\rm e}^{-\ii\,k\cdot x}\right) \right]\, . \nonumber \end{eqnarray} The matrix structure in Eq.~(\ref{phivertex2}) can be simplified using \begin{equation} \epsilon_{\gamma\delta}\, \left[ x\,\overline{\sigma}_\mu \right]^\delta_{\ \lambda}\, \left[ x\,\overline{k}\,U^\dagger\right]^\gamma_{\ m} = x^2\,\epsilon_{\gamma\delta}\,\left( U^\dagger\right)^\gamma_{\ m}\, \left[ k\,\overline{\sigma}_\mu \right]^\delta_{\ \lambda}\, , \end{equation} which follows from the transposition rules of the $\sigma$-matrices. Thus Eq.~(\ref{phivertex2}) can be rewritten as \begin{eqnarray} {\mathcal V}^{(t)}_{\mu \ m \lambda}\,(q,-k) &=& -\frac{\rho^{3/2}}{\pi}\, \int d^4x\,{\rm e}^{-\ii\,( q-k)\cdot x}\, \frac{1}{\left( x^2+\rho^2\right)^2} \label{phivertex3} \\ &\times & \epsilon_{\gamma\delta}\,\left( U^\dagger\right)^\gamma_{\ m}\, \left[ \left[ x\,\overline{\sigma}_\mu \right]^\delta_{\ \lambda} +\frac{1}{2}\,\rho^2\, \left[ k\,\overline{\sigma}_\mu \right]^\delta_{\ \lambda}\, \frac{1}{k\cdot x}\, \left( 1-{\rm e}^{-\ii\,k\cdot x}\right) \right]\, . \nonumber \end{eqnarray} The remaining $d^4x$ integration in Eq.~(\ref{phivertex3}) can be done with the help of the following formulae ($k^2=0$ is always understood), \begin{eqnarray} \label{int1} \lefteqn{ \int d^4x\,{\rm e}^{-\ii\,(q-k)\cdot x}\, \frac{x}{\left( x^2+\rho^2\right)^2} =} \\[1.6ex] \nonumber && -2\,\pi^2\,\ii\,\frac{q-k}{\left( q-k\right)^2}\, \rho\,\sqrt{\left( q-k\right)^2}\, K_1\left(\rho\,\sqrt{\left( q-k\right)^2}\right)\, , \\[2.4ex] \label{int2} \lefteqn{ \int d^4x\,{\rm e}^{-\ii\,(q-k)\cdot x}\, \frac{1}{\left( x^2+\rho^2\right)^2\,\left( k\cdot x\right)} = } \\[1.6ex] \nonumber && 2\,\pi^2\,\ii\, \frac{1}{q\cdot k}\,\frac{1}{\rho^2}\, \rho\,\sqrt{\left( q-k\right)^2}\, K_1\left(\rho\,\sqrt{\left( q-k\right)^2}\right)\, , \\[2.4ex] \label{int3} \lefteqn{ \int d^4x\,{\rm e}^{-\ii\,q\cdot x}\, \frac{1}{\left( x^2+\rho^2\right)^2\,\left( k\cdot x\right)} = } \\[1.6ex] \nonumber && 2\,\pi^2\,\ii\, \frac{1}{q\cdot k}\,\frac{1}{\rho^2}\, \rho\,\sqrt{q^2}\, K_1\left(\rho\,\sqrt{q^2}\right)\, . \end{eqnarray} By means of these basic integrals, we obtain finally for the vertex, \begin{eqnarray} \lefteqn{ {\mathcal V}^{(t)}_{\mu \ m \lambda}\,(q,-k) = 2\,\pi\,\ii\,\rho^{3/2}\, \epsilon_{\gamma\delta}\,\left( U^\dagger\right)^\gamma_{\ m} } \label{phivertex4} \\ &&\times \Biggl\{ \frac{\left[ \left(q-k\right) \,\overline{\sigma}_\mu \right]^\delta_{\ \lambda}} {(q-k)^2}\, \rho\,\sqrt{\left( q-k\right)^2}\, K_1\left(\rho\,\sqrt{\left( q-k\right)^2}\right) \nonumber \\ \mbox{}&&- \frac{1}{2}\, \frac{\left[ k\,\overline{\sigma}_\mu \right]^\delta_{\ \lambda}} {q\cdot k}\, \left[ \rho\,\sqrt{\left( q-k\right)^2}\, K_1\left(\rho\,\sqrt{\left( q-k\right)^2}\right) -\rho\,\sqrt{q^2}\, K_1\left(\rho\,\sqrt{q^2}\right) \right] \Biggr\}\, , \nonumber \end{eqnarray} in accordance with Eq.~(\ref{tvertex}). \vspace{10pt} \section*{Acknowledgements} We would like to acknowledge helpful discussions with V. Braun and V. Rubakov. \vspace{10pt}
proofpile-arXiv_065-483
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section*{Introduction} There has been recently an important activity in the study of {$N=2$} supersymmetric hierarchies (KP \cite{popo1,aratyn,dasb1,ghosh,dasp}, generalizations of KdV \cite{bokris,ikrim}, Two Bosons \cite{dasb2}, NLS \cite{kriso,krisoto,dasb3}, etc..). The most usual tools in this field are the algebra of $N=1$ pseudo-differential operators and Gelfand-Dickey type Poisson brackets \cite{geldik}. Although these systems have {$N=2$} supersymmetry, only for very few of them with very low number of fields is a formulation in extended superspace known. It is the purpose of this paper to partially fill this gap. The formalism which we shall present here partly originates from the article \cite{delma}. It turns out that in order to construct the Lax operators of {$N=2$} supersymmetric hierarchies, one should not use the whole algebra of {$N=2$} pseudo-differential operators, but rather the subalgebra of pseudo-differential operators preserving chirality. These operators were first considered in \cite{popo3}. They will be defined in section \ref{main}, where we also study the KP Lax equations and the two associated Hamiltonian structures. It turns out that the first (linear) bracket is associated with a non-antisymmetric $r$ matrix \cite{semenov}. Because of that, the second (quadratic) bracket is not of pure Gelfand-Dickey type. The main result of this paper is that we find two possibilities for this quadratic bracket. In fact, we show that there is an invertible map in the KP phase space which sends one of the quadratic Poisson structure into the other. However, this map does not preserve the Hamiltonians. In section \ref{reduc}, we study the possible reductions of the KP hierarchy by looking for Poisson subspaces in the phase space. These are different depending on the quadratic bracket which is used. Among these reductions, there are two different hierarchies with the {$N=2$} classical super-${\cal W}_n$ algebra \cite{lupope} as a hamiltonian structure. In particular, two of the three known {$N=2$} supersymmetric extensions of the KdV hierarchy \cite{Mathieu1} are found. They correspond to $a=-2$ and $a=4$ in the classification of Mathieu. These and some other examples are described in section \ref{examples}. Notice that from the known cases with a low number of fields \cite{Mathieu1,math2,popo2,yung1,Ivanov1,yung2}, one expects for any $n$ three hierarchies with super-${\cal W}_n$ as a hamiltonian structure. So our construction does not exhaust the possible cases. We also found two hierarchies which Poisson structure is the classical ``small" $N=4$ superconformal algebra. In one case the evolution equations are $N=4$ supersymmetric, while in the other they are only {$N=2$} supersymmetric. Finally, in section \ref{n1susy} we give the relation of our formulation with the usual formulation of the {$N=2$} supersymmetric KP Lax equations in $N=1$ superspace \cite{inami,dasb1,dasp}. \setcounter{equation}{0} \section{N=2 KP hierarchy \label{main}} \paragraph{{$N=2$} supersymmetry} We shall consider an {$N=2$} superspace with space coordinate $x$ and two Grassmann coordinates $\theta$, $\bar\theta$. We shall use the notation ${\underline x}$ for the triple of coordinates $(x,\theta,\bar\theta)$. The supersymmetric covariant derivatives are defined by \begin{equation} \partial\equiv{\partial\over\partial x},\,\,D={\partial\over\partial\theta} +\bar\theta\partial,\,\,\bar D={\partial\over\partial\bar\theta} +\theta\partial, D^2=\bar D^2=0,\,\,\{ D,\bar D\}=\partial \label{n2alg}\end{equation} Beside ordinary superfields $H({\underline x})$ depending arbitrarily on Grassmann coordinates, one can also define chiral superfields $\varphi({\underline x})$ satisfying $D\varphi =0$ and antichiral superfields $\bar\varphi({\underline x})$ satisfying $\bar D\bar\varphi =0$. We define the integration over the {$N=2$} superspace to be \begin{equation} \int d^3{\underline x}\, H(x,\theta,\bar\theta)= \int dx\bar DDH(x,\theta,\bar\theta) \vert_{\theta=\bar\theta=0}. \end{equation} The elements of the associative algebra of {$N=2$} pseudo-differential operators ($\Psi$DOs) are the operators \begin{equation} P = \sum_{i <M} ( a_{i} +b_i[D,\bar D]+\alpha_{i} D + \beta_{i} \overline{D} )\partial^{i} \label{pdo}\end{equation} where $a_{i}$, $b_{i}$ and $\alpha_{i}$, $\beta_{i}$ are respectively even and odd {$N=2$} superfields. However, this algebra is not very manageable. In particular, the set of strictly pseudo-differential operators ($M=0$ in\reff{pdo}) is not a proper subalgebra, but only a Lie subalgebra. Also, there are too many fields in these operators. We expect the phase space of the {$N=2$} KdV hierarchies to consist of the supercurrents of the {$N=2$} ${\cal W}_n$ algebras. In extended superspace, these supercurrents are bosonic superfields, and there is one such superfield for a given integer dimension. But in \reff{pdo}, each power of $\partial$ corresponds to four superfields, two even ones of integer dimension and two odd ones of half-integer dimension. It is thus clear that one has to restrict suitably the form of the {$N=2$} operators. It turns out that a possible restriction is to define the set $\cal C$ of pseudo-differential operators $L$ preserving chirality of the form \begin{footnote} {Operators of this type were first considered in \cite{popo3}} \end{footnote} \begin{equation} L=D{\cal L}\bar D,\,\,\,\,\,\,{\cal L}= \sum_{i <M} u_{i}\partial^{i} \label{cpdo}\end{equation} The coefficient functions $u_i$ are bosonic {$N=2$} superfields. These operators satisfy $DL=L\bar D=0$. The product of two chiral operators is again a chiral operator. The explicit product rule is easily worked out \begin{equation} LL'= D \left( {\cal L}\partial {\cal L'} +(D.{\cal L})(\bar{D}.{\cal L'}) \right) \bar{D}, \end{equation} where we have used the notation \begin{equation} (D.{\cal L})=\sum_{i <M}(Du_{i})\partial^{i}. \end{equation} Notice that $I=D \partial^{-1}\bar{D}$ is the unit of the algebra $\cal C$. We could have used as well the algebra $\bar{\cal C}$ of $\Psi$DOs satisfying $\bar D\bar L=\bar L D=0$. Notice that the product of an element in ${\cal C}\,\,$ by an element in $\bar{\cal C}$ vanishes. In fact ${\cal C}\,\,$ and $\bar{\cal C}$ are related by transposition, $L^t=-\bar D{\cal L}^tD\in \bar{\cal C}$. Although the transposition leads from ${\cal C}$ to $\bar{\cal C}$, there exists an anti-involution which acts inside ${\cal C}$. It is given by \begin{equation} \tau(L)=DL^t\partial^{-1}\bar D,\,\,\,\tau(L_{1}L_{2})=\tau(L_{2})\tau(L_{1}). \end{equation} Notice that it does not make sense in the algebra $\cal C$ to multiply a $\Psi$DO by a function. However, it is possible to multiply on the left by a chiral function $\phi$, $D\phi=0$ \begin{equation} \phi L= D\phi{\cal L}\bar D={\lambda}(\phi) L,\,\,\,{\lambda}(\phi)\equiv D\phi\partial^{-1}\bar D, \end{equation} and on the right by an antichiral function $\bar\phi$, $\bar D\bar\phi=0$ \begin{equation} L\bar\phi = D{\cal L}\bar\phi\bar D=L{\bar\lambda}(\bar\phi),\,\,\, {\bar\lambda}(\bar\phi)\equiv D\partial^{-1}\bar\phi\bar D. \end{equation} We define the residue of the pseudo-differential operator $L$ by $\mbox{\rm res} L=u_{-1}$ \cite{Mathieu1}. The residue of a commutator is a total derivative, $\mbox{\rm res}[L,L']=D\bar\omega+\bar D\omega$. The trace of $L$ is the integral of the residue \begin{equation} \mbox{\rm Tr}L=\int d^3{\underline x}\,\mbox{\rm res}L,\,\,\,\, \mbox{\rm Tr}[L,L']=0. \end{equation} $\cal C$ can be divided into two proper subalgebras ${\cal C} = {\cal C}_{+} \oplus {\cal C}_{-}$, where $L$ is in ${\cal C}_{+}$ if $\cal L$ is a differential operator and $L$ is in ${\cal C}_{-}$ if $\cal L$ is a strictly pseudo-differential operator ($M=0$ in \reff{cpdo}). We shall note \begin{equation} L=L_++L_-, \,\,\,L_+=D{\cal L}_+\bar D\in {\cal C}_+,\,\,\, L_-=D{\cal L}_-\bar D\in{\cal C}_-. \end{equation} Here an important difference with the usual bosonic and $N=1$ cases occurs. For any two $\Psi$DOs $L$ and $L'$ in $\cal C$ one has $\mbox{\rm Tr}(L_{-}L'_{-})= \int d^3{\underline x}\,\,\mbox{\rm res}(L)\,\,\mbox{\rm res}(L') \neq 0$. While ${\cal C}_{+}$ is an isotropic subalgebra, ${\cal C}_{-}$ is not. One important consequence of this fact is that if one defines the endomorphism $R$ of $\cal C$ by $R(L)={1\over 2}(L_+-L_-)$, then $R$ is a non-antisymmetric classical $r$ matrix, \begin{equation} \mbox{\rm Tr} (R(L)L'+LR(L'))=-\int d^3{\underline x}\,\mbox{\rm res} L\,\mbox{\rm res} L'. \end{equation} Notice that a non-antisymmetric $r$ matrix in the context of bosonic KP Lax equations first appeared in \cite{kuper}. \paragraph{KP equations} Let us now write the evolution equations of the {$N=2$} supersymmetric KP hierarchy. We consider operators $L=D{\cal L}\bar D$ in $\cal C$ of the form \begin{equation} {\cal L}=\partial^{n-1}+\sum_{i=1}^{\infty}V_i\partial^{n-i-1}. \label{KPop}\end{equation} $L$ has a unique $n$th root in $\cal C$ of the form \begin{equation} L^{1\over n}=D(1+\sum_{i=1}^{\infty}W_i\partial^{-i})\bar D, \end{equation} and we are led to consider the commuting flows (see \cite{popo3}) \begin{equation} {\partial\over\partial t_k}L=[(L^{k\over n})_+,L]=[R(L^{k\over n}),L]. \label{KPeq}\end{equation} There are symmetries of these equations which may be described as follows. Let us first introduce a chiral, Grassmann even superfield $\varphi$ which satisfies \begin{equation} {\partial\over\partial t_k}\varphi=(L^{k\over n})_+.\varphi \label{fieq}\end{equation} where the right-hand side is the chiral field obtained by acting with the differential operator $(L^{k\over n})_+$ on the field $\varphi$. Then the transformed operator \begin{equation} s(L)= \lambda(\varphi^{-1})L\lambda(\varphi) \label{simil}\end{equation} satisfies an evolution equation of the same form \reff{KPeq} as that of $L$. We may also consider an antichiral, Grassmann odd superfield $\bar\chi$ which satisfies \begin{equation} {\partial\over\partial t_k}\bar\chi=-(L^{k\over n})^t_+.\bar\chi \label{chieq}\end{equation} Then the transformed operator \begin{equation} \sigma(L)=(-1)^n\lambda((D\bar\chi)^{-1})\tau(L)\lambda(D\bar\chi) \label{chitra}\end{equation} satisfies an evolution equation of the same form \reff{KPeq} as that of $L$, with the direction of time reversed. \paragraph{Poisson brackets} The Lax equations \reff{KPeq} are bi-hamiltonian with respect to two compatible Poisson brackets which we now exhibit. Let $X$ be some $\Psi$DO in $\cal C$ with coefficients independent of the phase space fields $\{V_i\}$, then define the linear functional $l_{X}(L) = \mbox{\rm Tr}(LX)$. The generalization of the first Gelfand-Dickey bracket is obvious and reads \begin{equation} \{ l_{X},l_{Y} \}_{(1)} (L) = \mbox{\rm Tr} \left( L[X_{+},Y_{+}]-L[X_{-},Y_{-}] \right). \label{pb1}\end{equation} This is nothing but the linear bracket associated with the matrix $R$. Now we turn to the construction of the second bracket. It will turn out more complicated than the standard Gelfand-Dickey bracket because of the non-antisymmetry of the $r$ matrix. An analogous situation in the bosonic case is studied in \cite{Oevel}. We finally found two different possibilities. In order to write them down, we need to be able to separate the residue of a $\Psi$DO in $\cal C$ into a chiral and an antichiral part. For an arbitrary superfield $H({\underline x})$, we define \begin{equation} H=\Phi[H]+\bar\Phi[H],\,\,\, D\Phi[H]=0,\,\,\, \bar D\bar\Phi[H]=0. \end{equation} This is not a local operation in $\cal C$. An explicit form may be chosen as \begin{equation} \Phi[H]=D\bar D\int d^{3}{\underline x}'\Delta({\underline x}-{\underline x}')H({\underline x}'),\,\, \bar \Phi[H]=\bar DD\int d^{3}{\underline x}'\Delta({\underline x}-{\underline x}')H({\underline x}'), \label{chipro}\end{equation} where $\Delta$ is the distribution \begin{eqnarray} &\Delta({\underline x}-{\underline x}')= (\theta-\theta')(\bar\theta-\bar\theta')\epsilon(x-x'),&\label{distri}\\ &\partial\epsilon(x-x')=\delta(x-x'),\,\,\, \epsilon(x-x')=-\epsilon(x'-x). \nonumber\end{eqnarray} In the following, we shall use the short-hand notations $\Phi[\,\mbox{\rm res}[L,X]]=\Phi_X$, $\bar\Phi[\,\mbox{\rm res}[L,X]]=\bar\Phi_X$. In general, $\Phi_{X}$ will not satisfy the same boundary conditions as the phase space fields do. However, we noted earlier that in the case of a commutator, the residue is a total derivative, $\,\mbox{\rm res} [L,X]=D\bar\omega+\bar D\omega$. Here $\omega$ and $\bar\omega$ are differential polynomials in the fields. Then one easily shows that $\Phi_{X}=D\bar\omega+\alpha$, $\bar\Phi_{X}=\bar D\omega-\alpha$, where $\alpha$ is a constant reflecting the arbitrariness in the definition of $\Phi$, $\bar\Phi$. Up to this constant, $\Phi_{X}$ will respect the boundary conditions. We are now in a position to write the two possibilities for the second bracket as \begin{footnote} {The Poisson brackets (\ref{pb2p},\ref{pb2m}) may be put in the general form introduced in \cite{Maillet} \begin{equation} \{ l_{X},l_{Y} \}_{(2)}^{a,b} (L) =\mbox{\rm Tr} \left( LXa(LY)+XLb(LY)-LXc(YL)-XLd(YL) \right)\end{equation} However, the price to pay is that $a$, $b$, $c$, $d$ are non-local endomorphisms of $\cal C$. As an example, for the first quadratic bracket one finds \begin{equation} a(X)={1\over 2}(X_++\lambda(\Phi[\,\mbox{\rm res} X]))-{1\over 2} (X_--\lambda(\Phi[\,\mbox{\rm res} X])),\,\,\, b(X)=\bar\lambda(\bar\Phi[\,\mbox{\rm res} X]). \end{equation} One easily checks in particular that $a$ is a non-local antisymmetric $r$ matrix.} \end{footnote} \begin{equation} \{ l_{X},l_{Y} \}_{(2)}^a (L) =\mbox{\rm Tr} \left( LX(LY)_{+}-XL(YL)_{+}+ \Phi_Y LX+XL\bar\Phi_Y\right), \label{pb2p}\end{equation} and \begin{equation} \{ l_{X},l_{Y} \}_{(2)}^b (L) =\mbox{\rm Tr} \left( LX(LY)_{+}-XL(YL)_{+}+ \Phi_Y XL+LX\bar\Phi_Y\right). \label{pb2m}\end{equation} These expressions do not depend on the arbitrary constant $\alpha$. Checking the antisymmetry of the Poisson brackets and the Jacobi identity can be done with a little effort. As usual, the first bracket is a linearization of the two quadratic ones, that is to say \begin{equation} \{ l_{X},l_{Y} \}_{(2)}^{a,b} (L+zD\partial^{-1}\bar D) =\{ l_{X},l_{Y} \}_{(2)}^{a,b} (L) +z\{ l_{X},l_{Y} \}_{(1)} (L), \end{equation} and the linear bracket is compatible with each of the two quadratic brackets. Introducing the hamiltonians ${\cal H}_{k} = {n\over k}\mbox{\rm Tr}(L^{k\over n})$, the KP evolution equations \reff{KPeq} may be written as \begin{equation} \partial_{t_k} \left( l_{X}(L) \right) = \{ l_{X},{\cal H}_{k+n} \}_{(1)} (L) = \{ l_{X},{\cal H}_{k} \}_{(2)}^{a,b} (L) \end{equation} \paragraph{Poisson maps} Before turning to the study of the reductions of the KP hierarchies, let us exhibit some relations between the two quadratic brackets. We will use the invertible map in $\cal C$ \begin{equation} p(L)=\partial^{-1}\tau(L)=D\partial^{-1}L^t\partial^{-1}\bar D. \label{poimap}\end{equation} Then a straightforward calculation leads to \begin{equation} \{ l_{X}\circ p,l_{Y}\circ p\}_{(2)}^a=-\{ l_{X},l_{Y}\}_{(2)}^b\circ p, \end{equation} which shows that \reff{pb2p} and \reff{pb2m} are equivalent Poisson brackets. However there is no relation between the hamiltonians $\mbox{\rm Tr}(L^{k\over n})$ and $\mbox{\rm Tr}(p(L)^{k\over n-1})$. There is another relation between the two brackets, which involves the chiral superfield $\varphi$ satisfying the evolution equation \reff{fieq}. Let us introduce the linear functional $l_t=\int d^3{\underline x}(t\varphi)$, where $t({\underline x})$ is a Grassmann even superfield. We consider an enlarged phase space including $\varphi$, and extend the Poisson bracket \reff{pb2p} to this phase space by \begin{equation} \{ l_t,l_{Y} \}_{(2)}^a (L,\varphi)=\int d^3{\underline x} t((LY)_+.\varphi +\Phi_Y\varphi),\,\,\,\{ l_{t},l_{t'} \}_{(2)}^a=0. \end{equation} Then one finds \begin{equation} \{ l_{X}\circ s,l_{Y}\circ s\}_{(2)}^a=\{ l_{X},l_{Y}\}_{(2)}^b\circ s, \end{equation} where the transformation $s$ has been defined in \reff{simil}. Notice that the hamiltonians are invariant functions for the transformation $s$, $\mbox{\rm Tr}(L^{k\over n})=\mbox{\rm Tr}(s(L)^{k\over n})$. A last relation uses the antichiral superfield $\bar\chi$ satisfying the evolution \reff{chieq}. Let us introduce the linear functional $l_{\bar t}=\int d^3{\underline x}(\bar t\bar\chi)$, where $\bar t({\underline x})$ is a Grassmann odd superfield. We consider an enlarged phase space including $\bar\chi$, and extend the Poisson bracket \reff{pb2p} to this phase space by \begin{eqnarray} &\{ l_{\bar t},l_{Y} \}_{(2)}^a (L,\bar\chi)=\int d^3{\underline x} {\bar t}(-(LY)^t_+.\bar\chi +\Phi_Y\bar\chi),&\\ & \{ l_{\bar t_1},l_{\bar t_1} \}_{(2)}^a=-2\int d^{3}{\underline x}\bar t_1\bar\chi\bar\Phi[\bar t_2\bar\chi],& \end{eqnarray} where $\Phi$, $\bar\Phi$ are defined in equations (\ref{chipro},\ref{distri}). Notice that this is a non-local Poisson bracket. One finds \begin{equation} \{ l_{X}\circ\sigma,l_{Y}\circ\sigma\}_{(2)}^a=-\{ l_{X},l_{Y}\}_{(2)}^b\circ\sigma, \end{equation} where the transformation $\sigma$ has been defined in \reff{chitra}. \setcounter{equation}{0} \section{Reductions of the KP hierarchy \label{reduc}} In order to obtain consistent reductions of the KP hierarchy, we need to find Poisson submanifolds of the KP phase space. Considering first the quadratic bracket \reff{pb2p}, we rewrite it as \begin{eqnarray} &\{ l_{X},l_{Y} \}_{(2)}^a (L) =\mbox{\rm Tr} X\xi_{l_Y}^a,&\nonumber\\ &\xi_{l_Y}^a=(LY)_{+}L-L(YL)_{+}+\Phi_Y L+L\bar\Phi_Y.& \label{hvf}\end{eqnarray} $\xi_{l_Y}^a$ is the hamiltonian vector field associated with the function $l_Y$. One easily checks that if L has the form \reff{KPop}, then for any $Y$, $\xi_{l_Y}^a$ has the form $D(\sum_{i<n-1}\xi_i\partial^i)\bar D$. It is obvious from \reff{hvf} that for any $Y$, if $L$ is in ${\cal C}_+$, then $\xi_{l_Y}^a$ is also in ${\cal C}_+$. This means that the constraint \begin{equation}L=L_+ \label{kdv}\end{equation} defines a Poisson submanifold. The hierarchies obtained in this way are the {$N=2$} supersymmetric KdV hierarchies studied by Inami and Kanno \cite{inami}, and the Lax operators \reff{kdv} already appeared in \cite{popo3}. The lowest order cases will be presented in the next section. Another possible reduction is to take $L$ of the form \begin{equation} L=L_++D\,\varphi\partial^{-1}\bar\varphi\bar D,\,\,\,\,\, D\varphi=\bar D\bar\varphi=0. \label{nls}\end{equation} where $\varphi$ and $\bar\varphi$ are Grassmann even or odd chiral superfields. With $L$ of the form \reff{nls} and $Y$ arbitrary, one finds \begin{equation} (\xi_{l_Y}^a)_-=D((LY)_+.\varphi+\Phi_Y\varphi)\partial^{-1}\bar\varphi +\varphi\partial^{-1}(-(YL)_+^t.\bar\varphi+\bar\Phi_Y\bar\varphi))\bar D, \end{equation} Noticing that $(LY)_+.\varphi$ is a chiral superfield and $(YL)_+^t.\bar\varphi$ an antichiral superfield, it is easily checked that $\xi_{l_Y}^a$ is indeed tangent to the submanifold defined by the constraints \reff{nls}. It is possible to consider an enlarged phase space which coordinates are the fields in $L$ and $\varphi$, $\bar\varphi$. Let us introduce the linear functionals \begin{equation}l_t=\int d^3{\underline x}(\varphi t),\,\, l_{\bar t}=\int d^3{\underline x}(\bar t\bar\varphi), \end{equation} where $ t$ and $\bar t$ are general superfields, of the same Grassmann parity as $\varphi$ and $\bar\varphi$. In this enlarged phase space, the second Poisson bracket, in the case when $\varphi$ and $\bar\varphi$ are Grassmann even, is defined by \reff{pb2p} and \begin{eqnarray} &\{ l_t,l_{Y} \}_{(2)}^a (L,\varphi,\bar\varphi)=\int d^3{\underline x} ((LY)_+.\varphi +\Phi_Y\varphi)t, &\label{lfi}\\ &\{l_{\bar t},l_{Y} \}_{(2)}^a (L,\varphi,\bar\varphi)=\int d^3 {\underline x}\,\bar t(-(YL)_+^t.\bar\varphi +\bar\Phi_Y\bar\varphi),& \nonumber\end{eqnarray} and \begin{eqnarray} &\{ l_t,l_{\bar t} \}_{(2)}^a (L,\varphi,\bar\varphi)= \int d^3{\underline x}\, ({ L}_+.\bar t)t,&\label{bose}\\& \{ l_{t_1},l_{t_2} \}_{(2)}^a=0,\,\,\,\, \{ l_{\bar t_1},l_{\bar t_2} \}_{(2)}^a=0.& \nonumber\end{eqnarray} In the case when $\varphi$ and $\bar\varphi$ are Grassmann odd, the last two lines should be modified to \begin{eqnarray} &\{ l_t,l_{\bar t} \}_{(2)}^a (L,\varphi,\bar\varphi)= \int d^3{\underline x} (({ L}_+.\bar t)t-2\varphi t\Phi[\bar t\bar\varphi]), \label{fermi}&\\& \{ l_{t_1},l_{t_2} \}_{(2)}^a=2\int d^3{\underline x}\,\varphi t_1 \Phi[\varphi t_2],\,\,\,\, \{ l_{\bar t_1},l_{\bar t_2} \}_{(2)}^a=-2\int d^3{\underline x} \,\bar t_1\bar\varphi \bar\Phi[\bar t_2\bar\varphi],& \nonumber\end{eqnarray} where the applications $\Phi$ and $\bar\Phi$ have been defined in \reff{chipro}. The lowest order case is $L=D(1+\varphi\partial^{-1}\bar\varphi)\bar D$. Then if $\varphi$ and $\bar\varphi$ are odd, the equation ${d\over dt}L=[L^2_+,L]$ is the {$N=2$} supersymmetric extension of the NLS equation \cite{roelo}. The next-to-lowest order case is $L=D(\partial +H+\varphi\partial^{-1}\bar\varphi)\bar D$. If $\varphi$ and $\bar\varphi$ are even, the hamiltonian structure \reff{pb2p} reduces in this case to the classical version of the ``small'' $N=4$ superconformal algebra. Although the Poisson algebra contains $4$ supersymmetry generators, the evolution equations \reff{KPeq} have only {$N=2$} supersymmetry. This case was first obtained by another method which will be given, as part of a detailed study, in \cite{dgi}. We now turn to the second quadratic bracket \reff{pb2m}. We rewrite it as \begin{eqnarray} &\{ l_{X},l_{Y} \}_{(2)}^b (L) =\mbox{\rm Tr} X\xi_{l_Y}^b,&\nonumber\\ &\xi_{l_Y}^b=(LY)_{+}L-L(YL)_{+}+L{\lambda}(\Phi_Y)+{\bar\lambda}(\bar\Phi_Y)L.& \label{hvfm}\end{eqnarray} It is easily seen that neither the condition \reff{kdv}, nor the more complicated condition \reff{nls} are admissible reductions in this case. The easiest way to find Poisson subspaces for the bracket \reff{pb2m} is to apply the map \reff{poimap} to the Poisson subspaces of the first quadratic bracket. From \reff{kdv}, we are then lead to the restriction: \begin{equation} L=L_++D\bar D\partial^{-1}H\partial^{-1}D\bar D \label{a4}\end{equation} With $L$ of the form \reff{a4} and $Y$ arbitrary, one finds \begin{equation} (\xi_{l_Y}^b)_-=D\bar D\partial^{-1}((LY)_+.H-(YL)_+^t.H+\,\mbox{\rm res}[L,Y]H) \partial^{-1}D\bar D, \end{equation} which directly shows that condition \reff{a4} defines a Poisson submanifold for the Poisson bracket \reff{pb2m}. It turns out that \reff{a4} also defines a Poisson submanifold for the linear Poisson bracket \reff{pb1}. To show this we rewrite the linear bracket as \begin{equation}\{ l_{X},l_{Y} \}_{(1)} (L) =\mbox{\rm Tr} X\eta_{l_Y},\,\,\, \eta_{l_Y}=[L,Y]_+-[L,Y_+]+{\lambda}(\Phi_Y)+{\bar\lambda}(\bar\Phi_Y). \end{equation} With $L$ of the form \reff{a4} and $Y$ arbitrary, one finds \begin{equation} (\eta_{l_Y})_-=D\bar D\partial^{-1}((Y_+-Y_+^t).H+\,\mbox{\rm res}[L,Y])\partial^{-1}D\bar D. \end{equation} Thus the reduced hierarchies defined by condition \reff{a4} are bi-hamiltonian. The lowest order cases will be studied in the next section. Notice that the transformation \reff{simil} maps the systems satisfying the condition \reff{nls} with Grassmann even fields $\varphi$ and $\bar\varphi$ into systems satisfying condition \reff{a4} with \begin{equation} H= \varphi\bar\varphi+\varphi^{-1}L_+.\varphi. \end{equation} Analogously, the transformation \reff{chitra} maps the systems satisfying the condition \reff{nls} with Grassmann odd fields $\varphi$ and $\bar\varphi$ into systems satisfying condition \reff{a4} with \begin{equation} H= (-1)^n\left(\bar\varphi\varphi+(D\bar\varphi)^{-1}D(L_+^t.\bar\varphi)\right). \end{equation} Such transformations may be found in \cite{kriso,bokris}. Finally we may consider the image of the Poisson subspace defined by \reff{nls} under the map $p$. One finds the condition \begin{equation} L=L_++D\bar D\partial^{-1}(H+\bar\varphi\partial^{-1}\varphi)\partial^{-1}D\bar D. \label{n4}\end{equation} The lowest order case is when $L_{+}=D\bar D$. The hamiltonian structure \reff{pb2m} reduces in this case to the classical version of the ``small'' $N=4$ superconformal algebra. The equation ${d\over dt}L=[(L^{3})_{+}, L]$ becomes, after suitable redefinitions, the $N=4$ supersymmetric extension of the KdV equation derived in \cite{delivan} and written in {$N=2$} superspace in \cite{dik}. One can again consider an enlarged phase space which coordinates are the fields in $L$ and $\varphi$, $\bar\varphi$. The second quadratic bracket in this phase space is easily obtained from the first one by applying the map $p$ to the first quadratic bracket. $p$ acts as the identity on $\varphi$ and $\bar\varphi$. As a consequence the Poisson brackets \reff{bose} and \reff{fermi} keep the same form, whereas \reff{lfi} should be modified to \begin{eqnarray} &\{ l_t,l_{Y} \}_{(2)}^b (L,\varphi,\bar\varphi)=\int d^3{\underline x} (\,\mbox{\rm res}\left(\tau((YL)_+)\lambda(\varphi)\right) +\Phi_Y\varphi)t, &\\ &\{l_{\bar t},l_{Y} \}_{(2)}^b (L,\varphi,\bar\varphi)=\int d^3 {\underline x}\,\bar t(-\,\mbox{\rm res}\left(\bar\lambda(\bar\varphi)\partial^{-1}\tau((LY)_+)\partial\right) +\bar\Phi_Y\bar\varphi).& \nonumber\end{eqnarray} \setcounter{equation}{0} \section{Examples and comparison with other works \label{examples}} This paragraph is devoted to the presentation of the simplest integrable equations obtained using our formalism. Considering first the condition \reff{kdv}, the simplest example is the lax operator $L = D (\partial + W) \bar{D}$. Then the evolution equation \begin{equation} {d\over dt}L=[L^{3\over 2}_+,L], \end{equation} leads to the equation \begin{equation} 8\partial_{t} W = 2W_{xxx}+6\left( (DW)(\overline{D}W) \right)_{x}- \left( W^3 \right)_{x}, \end{equation} which coincide after the redefinition $W=2i\Phi$ with the $a=-2$ {$N=2$} extension of the KdV equation in the classification of Mathieu \cite{Mathieu1,math2}. The Lax operator given in \cite{Mathieu1} may be obtained from $L$ in the following way. Let us consider the operator \begin{equation} L_{-2} = L+L^t=\partial^2+W[D,\bar D]+(DW)\bar D-(\bar D W)D. \label{eqam2}\end{equation} $L$ is in $\cal C$ and $L^t$ is in $\bar{\cal C}$. If we remember that the product of an element in $\cal C$ and an element in $\bar{\cal C}$ always vanishes, we immediately get that a square root of $L_2$ with highest derivative term equal to $\partial$ is $(L_{-2})^{1\over 2}=L^{1\over 2}-(L^{1\over 2})^t$. From this we deduce the relation $(L_{-2})^{3\over 2}=L^{3\over 2} -(L^{3\over 2})^t$. As a consequence $L_{-2}$ satisfies the evolution equation \begin{equation}{d\over dt}L_{-2}=[L^{3\over 2}_+,L]+([L^{3\over 2}_+,L])^t =[(L_{-2})^{3\over 2},L_{-2}],\end{equation} which is thus an equivalent Lax representation for equation \reff{eqam2}. As the next example, we consider the Lax operator \begin{equation} L=D(\partial^2+V\partial+W)\bar D \end{equation} Then the evolution equation ${d\over dt}L=[L^{2/3}_{+},L]$ should coincide, after suitable redefinitions, with one of the three {$N=2$} supersymmetric extensions of the Boussinesq equations derived in \cite{Ivanov1}. Indeed one can check that the Lax operator they give for the $\alpha = -1/2$ equation may be written as $L^{(1)}=L+\bar D\partial^2 D$. Then one easily obtains $(L^{(1)})^{2\over 3}=L^{2\over 3}+\bar D\partial D$, and the evolution equation for $L^{(1)}$ is easily deduced from that of $L$ \begin{equation} {d\over dt}L^{(1)}={d\over dt}L=[(L^{(1)})^{2\over 3},L^{(1)}]. \end{equation} Turning now to condition \reff{a4}, the lowest order case corresponds to the Lax operator $L=D\bar D+D\bar D\partial^{-1}W\partial^{-1}D\bar D$. Then the equation ${d\over dt}L=[(L^{3})_{+}, L]$ becomes, after suitable redefinitions, the {$N=2$} supersymmetric extension of the KdV equation with parameter $a=4$, \begin{equation} \partial_{t}W = W_{xxx} + \frac{3}{2}\left( [ D ,\overline{D} ] W^2 \right)_{x} - 3\left( (DW)(\overline{D}W) \right)_{x} + (W^3)_{x} \end{equation} Notice that, all integer powers of $L$ define conserved charges in this case (an alternative Lax operator with the same property was derived in \cite{krisoto}). The last example that we shall study is the Lax operator \begin{equation} L = D \left( \partial + V \right) \overline{D} + D\overline{D}\partial^{-1}W\partial^{-1}D\overline{D}. \end{equation} Then the equation \begin{equation} \partial_{2} L = [L_{+},L] \end{equation} explicitely reads \begin{eqnarray} \partial_{2} V &=& 2 W_{x}\\ \partial_{2} W &=& [ D, \overline{D} ]W_{x} +VW_{x} +(DV)( \overline{D}W)+( \overline{D}V)(DW). \end{eqnarray} This equation is identical, up to a rescaling of time, to the {$N=2$} supersymmetric extension of the Boussinesq equation with parameter $\alpha = -2$ derived in \cite{Ivanov1}. \setcounter{equation}{0} \section{From $N=2$ to $N=1$ superspace \label{n1susy}} $N=2$ extensions of the KP and KdV hierarchies have been studied in several articles \cite{inami,ghosh,dasb1,dasp} using an $N=1$ superspace formalism. In this section we wish to relate the KP hierarchies that we described in section \ref{main} to those given in the litterature. The first step will be to relate our $N=2$ algebra ${\cal C}$ of pseudo-differential operators to the $N=1$ algebra of pseudo-differential operators. An operator $L=D{\cal L}\bar D$ in ${\cal C}$ should be considered as acting on a chiral object $\Psi$, $D\Psi$=0, and this action writes \begin{equation} L .\Psi=D{\cal L}\bar D .\Psi={\cal L}\partial .\Psi+(D .{\cal L})\bar D .\Psi.\label{R1} \end{equation} We shall use the following combinations of the chiral derivatives \begin{equation} D_1=D+\bar D,\,\, D_2=-D+\bar D,\,\, D_1^2=-D_2^2=\partial,\,\, \{ D_1,D_2\}=0.\end{equation} Then the action of $L$ on $\Psi$ is \begin{equation} L .\Psi=({\cal L}\partial+(D .{\cal L})D_1) .\Psi. \end{equation} We then choose to associate to the $N=2$ pseudo-differential operator $L$ the $N=1$ pseudo-differential operator $\underline L$ given by \begin{equation} {\underline L}={\cal L}\vert_{\theta_2=0}\partial+ (D .{\cal L})\vert_{\theta_2=0}D_1. \end{equation} It is easily checked that this correspondence respects the product, $\underline{LL'}=\underline{L}\,\,\underline{L'}$. It also has the property \begin{equation} \underline{L_+}={\underline L}_{>0}. \end{equation} That is to say that the image of an $N=2$ differential operator is a strictly differential $N=1$ operator, without the non-derivative term. Notice also the useful relations \begin{eqnarray} & \,\mbox{\rm res} (\underline{L})=(D. {\,\mbox{\rm res}}(L))\vert_{\theta_2=0},\,\, &\\ &\mbox{\rm Tr}(L)=\mbox{\rm Tr}(\underline{L})\equiv \int d^2{\underline x} {\,\mbox{\rm res}}(\underline{L}),\,\,\,\int d^2{\underline x}\equiv\int dxd\theta_1& \end{eqnarray} where the residue of the operator $\underline L$ is the coefficient of $D_1^{-1}\equiv D_1\partial^{-1}$. From now on, all expressions will be written in $N=1$ superspace, and we drop the index of $D_1$ and $\theta_1$. The KP hierarchy described in section \ref{main} may be described in $N=1$ superspace as follows. We consider an operator $\underline L$ of the form \begin{equation} \underline{L}=D^{2n}+\sum_{p=1}^\infty w_pD^{2n-p-1} \end{equation} and consider evolution equations \begin{equation} {\partial\over\partial t_k}\underline{L}=[\underline{L}^{k\over n}_{>0},L] \label{R9}\end{equation} This is nothing but the non-standard supersymmetric KP hierarchy described in \cite{ghosh,dasb1}. The evolution equations (\ref{R9}) admit the conserved quantities $H_p=\mbox{\rm Tr}(\underline{L}^{p\over n})$, and they are bi-hamiltonian. The first Poisson bracket is easily deduced from its $N=2$ counterpart (\ref{pb1}). With $l_{\underline{X}}=\mbox{\rm Tr}(\underline{L}\,\,\underline{X})$, we have \begin{equation} \{ l_{\underline{X}},l_{\underline{Y}} \}_1=\mbox{\rm Tr} L([\underline{X}_{>0},\underline{Y}_{>0}]- [\underline{X}_{\leq 0},\underline{Y}_{\leq 0}]) \end{equation} As in the $N=2$ formalism, this is a standard bracket associated with a non-antisymmetric $r$ matrix. As a consequence, the two quadratic brackets deduced from (\ref{pb2p}) and \reff{pb2m} are quite complicated. They involve the quantity $\psi_{\underline{X}}$ defined up to a constant by $D\psi_{\underline{X}}= {\,\mbox{\rm res}}[{\underline{L}}\, ,{\underline{X}}]$. The first one is \begin{eqnarray} &\{ l_{\underline{X}},l_{\underline{Y}} \}_2^a(L)=\mbox{\rm Tr}(\underline{L}\,\,\underline{X}(\underline{L}\,\,\underline{Y})_+ -\underline{X}\,\,\underline{L}(\underline{Y}\,\,\underline{L})_+) +\int d^2{\underline x} (-\psi_{\underline{Y}}\, {\,\mbox{\rm res}}[{\underline{L}},{\underline{X}}]\nonumber&\\& +{\,\mbox{\rm res}}[{\underline{L}}\, ,{\underline{Y}}]\, {\,\mbox{\rm res}}(\underline{X}\,\,\underline{L}\,D^{-1}) - {\,\mbox{\rm res}}[{\underline{L}}\, ,{\underline{X}}]\, {\,\mbox{\rm res}}(\underline{Y}\,\,\underline{L}\,D^{-1}) ).& \end{eqnarray} The Poisson bracket \reff{pb2m} becomes \begin{eqnarray} &\{ l_{\underline{X}},l_{\underline{Y}} \}_2^b(L)=\mbox{\rm Tr}(\underline{L}\,\,\underline{X}(\underline{L}\,\,\underline{Y})_+ -\underline{X}\,\,\underline{L}(\underline{Y}\,\,\underline{L})_+) +\int d^2{\underline x} (\psi_{\underline{Y}}\, {\,\mbox{\rm res}}[{\underline{L}},{\underline{X}}] \nonumber&\\& +{\,\mbox{\rm res}}[{\underline{L}}\, ,{\underline{Y}}]\, {\,\mbox{\rm res}}(\underline{L}\,\,\underline{X}\,D^{-1}) - {\,\mbox{\rm res}}[{\underline{L}}\, ,{\underline{X}}]\, {\,\mbox{\rm res}}(\underline{L}\,\,\underline{Y}\,D^{-1}) ),& \end{eqnarray} and already appeared in \cite{dasp}. It is not a difficult task to obtain the $N=1$ restrictions which correspond to the {$N=2$} conditions (\ref{kdv},\ref{nls},\ref{a4},\ref{n4}). Some of the lax operators obtained in this way are already known, in particular those satisfying \reff{kdv} from \cite{inami} and the lowest order operator coming from \reff{nls} with odd $\varphi$ and $\bar\varphi$, which is the super-NLS Lax operator obtained in \cite{dasb1}. \setcounter{equation}{0} \section{Conclusion} An easy generalization of the hierarchies presented in this article would be to consider multi-components KP hierarchies, that is to say replace the fields $\varphi$ and $\bar\varphi$ in \reff{nls} and \reff{n4} by a set of $n+m$ fields $\varphi_i$ and $\bar\varphi_i$, $n$ of them being Grassmann even and the other $m$ being Grassmann odd. For the lowest order case of equation \reff{nls}, such a generalization has been considered in \cite{bokris}. The Lax representation that we propose for such hierarchies has the advantage that one does not need to modify the definition of the residue. For the next to lowest order case of equation \reff{nls}, and the lowest order case of equation \reff{n4}, it should be possible to obtain in this way hierarchies based on $\cal W$-superalgebras with an arbitrary number of supersymmetry charges. Little is known about the matrix Lax formulation of the hierarchies presented here. In the case of operators satisfying condition \reff{kdv}, such a matrix Lax formulation was constructed in $N=1$ superspace by Inami and Kanno \cite{inami,ina1}. It involves the loop superalgebra based on $sl(n\vert n)$. What we know about the matrix Lax formulation in {$N=2$} superspace for hierarchies based on Lax operators satisfying conditions \reff{kdv} or \reff{nls} will be reported elsewhere. Notice that we obtained the form \reff{kdv} of the scalar Lax operators from a matrix Lax representation, and only later became aware of reference \cite{popo3} where these operators also appear.
proofpile-arXiv_065-484
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Motivations} The CLEO Collaboration has very recently presented measurements of the branching ratios for $B\to\rho\ell\bar\nu$ and $B\to\pi\ell\bar\nu$ decays ($\ell{=}e,\mu$)~\cite{cleo} and has announced its intention to measure the corresponding differential decay rates. These various measurements represent an excellent opportunity to determine the poorly known CKM matrix element $|V_{ub}|$. Such determinations require understanding the non-perturbative, strong-interaction corrections to the elementary $b-u-W$ coupling contained in the matrix elements of the weak currents $V^\mu{=}\bar u\gamma^\mu b$ and $A^\mu{=}\bar u\gamma^\mu\gamma^5 b$ between $B$ and $\pi$ or $\rho$ meson states. It is to calculate these matrix elements that we resort to the lattice. $heavy\to light$ quark decays, such as the ones that concern us here, are also interesting because they enable one to test heavy-quark symmetry (HQS). For these decays, HQS is weaker than for $heavy\to heavy$ quark decays: it only applies in a limited region around the zero recoil point $q^2{=}q^2_{max}$, where $q$ is the four-momentum transferred to the leptons, and imposes no normalization condition on the relevant form factors at $q^2_{max}$. Nevertheless, because both the mass and the spin of the heavy quark can be varied in lattice calculations, the deviations from the heavy-quark limit due to finite heavy-quark mass and spin effects can be measured. \section{Limitations} Current day lattice calculations, with lattice spacings on the order of $3\mbox{ GeV}^{-1}$, do not permit one to simulate $b\to light$ quark decays over their full kinematical range. The problem is that the energies and momenta of the particles involved, whose orders of magnitude are set by the $b$ quark mass ($m_b\simeq 5\mbox{ GeV}$), are large on the scale of the cutoff in much of phase space. To limit these energies in relativsitc lattice quark calculations, one performs the simulation with heavy-quark mass values $m_Q$ around that of the charm ($m_c\simeq 1.5\mbox{ GeV}$), where discretization errors remain under control. Then one extrapolates the results up to $m_b$ by fitting heavy-quark scaling relations (HQSR) with power corrections to the lattice results (see \sec{hqex}). Another approach is to work with discretized versions of effective theories such as Non-Relativistic QCD (NRQCD) or Heavy-Quark Effective Theory (HQET) in which the mass of the heavy quark is factored out of the dynamics. All approaches, however, are constrained to relatively small momentum transfers because of the limited applicability of HQS and because of momentum-dependent discretization errors. So one can only reconstruct the $q^2$ dependence of the relevant form factors in a limited region around $q^2_{max}$ and one is left with the problem of extrapolating these results to smaller $q^2$. $heavy\to light$ quark decays are difficult in any theoretical approach. Indeed, they require understanding the underlying QCD dynamics over a large range of momentum transfers from $q^2_{max}{=}26.4\mbox{ GeV}^2 (20.3\mbox{ GeV}^2)$ for semileptonic $B\to\pi$ ($B\to\rho$) decays, where the final state hadron is at rest in the frame of the $B$ meson, to $q^2{=}0$ where it recoils very strongly. \section{$\bar B^0\to\rho^+\ell^-\bar\nu$ and a Model-Independent Determination of $|V_{ub}|$ \protect\footnote[1] {The results presented here were obtained on a $24^3\times 48$ lattice at $\beta=6.2$ from 60 quenched configurations, using an $\ord{a}$-improved SW action for the quarks. (Please see \protect\cite{JoF96} for details.)} } \label{sec:btorho} One solution to the problem of the limited kinematical range of the lattice results is to ignore the problem or rather rely on the ingenuity of experimental groups to provide measurements of partial rates in the region where lattice results are available. Combined with lattice results for $B\to\rho\ell\bar\nu$ decays, such experimental measurements will enable a model-independent determination of $|V_{ub}|$~\cite{btorho}. Rates should be sufficient since our lattice results span a range of $q^2$ from $\sim 14.4\mbox{ GeV}^2$ to $q^2_{max}$ over which the partially integrated is $4.6\er{4}{3}|V_{ub}|^2 ps^{-1}$. This represents approximatively $1/3$ of the total rate obtained from light-cone sumrules (LCSR) in \re{PBa96}, whose results at large $q^2$ agree well with ours. \subsection{Form Factors and Heavy-Quark Extrapolation} \label{hqex} To describe $\bar B^0\to\rho^+\ell^-\bar\nu$ decays, we must evaluate the matrix element $\langle \rho^+(p',\eta)|V^\mu-A^\mu|\bar B^0(p)\rangle$, traditionally decomposed in terms of four form factors $A_1$, $A_2$, $A$ and $V$ which are functions of $q^2$, where $q{=}p-p'$. We calculate this matrix element for four values of the heavy-quark mass around that of the charm. Then, to obtain $A_1$ at the scale of the $B$ meson we fit, to the lattice results, the HQSR \begin{eqnarray} A_1(\omega,M)\alpha_s(M)^{2/\beta_o}\sqrt{M}&= c(\omega)\big(1+ \frac{d(\omega)}{M}\nonumber\\ &+\ord{\Lambda^2/M^2}\big) ,\label{eq:hqsr} \end{eqnarray} where $M$ is the mass of the decaying meson, $\beta_0{=}11-2n_f/3$ and $\Lambda$ is an energy characteristic of the light degrees of freedom. This scaling relation holds for $\omega{=}(M^2+m^2-q^2)/2Mm$ close to 1 and the fit parameters $c$ and $d$ are independent of $M$. Here $m$ is the mass of the final state meson. Once $c$ and $d$ are fixed, it is trivial to obtain $A_1(M{=}m_B)$ at the corresponding value of $\omega$. Furthermore, $d$ determines the size of corrections to the heavy-quark limit. Repeating this procedure for many values of $\omega$, one obtains the $q^2$ dependence of the desired form factor around $q^2_{max}$. The resulting $A_1(q^2,m_B)$ is plotted together with the LCSR result of \re{PBa96} and the lattice results of APE~\cite{ape} and ELC~\cite{elc}. Agreement amongst these four calculations is excellent as it for $A_2$ and $V$~\cite{PBa96,JoF96}, which are obtained in an entirely analogous way. \begin{figure}[tb] \setlength{\epsfxsize}{60mm}\epsfbox[93 290 460 550]{a1.eps} \unit=0.8\hsize \point 0.90 -0.00 {\large{$q^2(\mbox{GeV}^2)$}} \point 0.30 0.55 {\Large{$A_1(q^2)$}} \caption{$A_1$ vs. $q^2$ from UKQCD (crosses), APE\protect\cite{ape} (right-hand dot), ELC\protect\cite{elc} (left-hand dot) and LCSR\protect\cite{PBa96} (curve). (Adapted from \protect \re{PBa96}.)} \label{fig:comp} \end{figure} \subsection{Rates} Having determined $A_1$, $A_2$ and $V$, we can compute $(1/|V_{ub}|^2)$$d\Gamma/dq^2$. Our results are plotted in \fig{fig:rate} (squares). In the region of $q^2$ accessed, we can legitimately expand, around $q^2_{max}$, the helicity amplitudes that appear in the rate. Thus we fit, to the lattice points, the parametrization \begin{eqnarray} \frac{10^{12}}{|V_{ub}|^2} \frac{d\Gamma}{dq^2}&\simeq&\frac{G^2_F}{192\pi^3m_B^3} q^2\lambda(q^2)^{1/2} \nonumber\\ &\times& a^2\left(1+b(q^2-q^2_{max})\right) \label{eq:rate} \end{eqnarray} where $\lambda(q^2){=}(M^2+m^2-q^2)- 4M^2m^2$. We find $a{=}4.6\er{4}{3}\pm 0.6 \mbox{ GeV}$ and $b{=}(-8\er{4}{6}) 10^{-2} \mbox{ GeV}^{-2}$ where the second error on $a$ is systematic, all other errors being statistical. With $a$ and $b$ determined, the only unknown in \eq{eq:rate} is $|V_{ub}|$. Therefore, a fit of the parametrization of \eq{eq:rate} to an experimental measurement of the differential decay rate around $q^2_{max}$ determines $|V_{ub}|$. In this determination, $a$ plays the role of ${\cal F}(1)$ in the extraction of $|V_{cb}|$ from semileptonic $B\to D$ or $D^*$ decays \cite{MNe94} and $b$ the role of ${\cal F}'(1)$. The difference, here, is that $a$ is not determined by HQS up to small radiative and power corrections. It is a genuinely non-perturbative quantity. Another way of determining $|V_{ub}|$ from the lattice results is to compare partially integrated rates from $q^2\ge 14\mbox{ GeV}^2$ to $q^2_{max}$ given by \eq{eq:rate} to the corresponding experimental measurements. Both these methods yield $|V_{ub}|$ with approximatively 10\% statistical and 12\% theoretical uncertainties. \begin{figure}[tb] \setlength{\epsfxsize}{60mm}\epsfbox[30 100 500 530]{vub.ps} \caption{The data points are our lattice results and the solid curve, the fit to \protect\eq{eq:rate}.} \label{fig:rate} \end{figure} \subsection{A Test of HQS} In \fig{fig:sem_over_rad} we compare semileptonic $B\to\rho$ form factors with those governing the short distance contribution to radiative $B\to K^*\gamma$ decays for which the relevant hadronic matrix element is $\langle K^*(p',\eta)|\bar s\sigma^{\mu\nu}q^\nu b_R|B(p)\rangle$, with $q{=}p-p'$. This matrix element is traditionally decomposed in terms of three form factors, $T_1$, $T_2$ and $T_3$. The comparison is made for three initial meson masses: $M{=}m_D$, $M{=}m_B$ and $M\to\infty$. For identical final-state vector mesons (in \fig{fig:sem_over_rad} all light-quarks involved have the same mass, slightly larger than that of the strange), HQS predicts \begin{equation} V(q^2)=2T_1(q^2),\quad A_1=2iT_2(q^2) \ ,\label{eq:sor} \end{equation} for $q^2$ around $q^2_{max}$ or, equivalently, $\omega$ close to 1. \begin{figure}[tb] \setlength{\epsfxsize}{30mm}\epsfbox[170 290 285 690]{sem_over_rad.ps} \caption{Ratios $V/2T_1$ and $A_1/2iT_2$ for 5 values of $\omega$ and three initial meson masses. The solid lines are the HQS predictions.} \label{fig:sem_over_rad} \end{figure} While $V/2T_1$ displays large $1/M$ corrections at the $D$ and even $B$ meson scale, $A_1/2iT_2$ exhibits no such corrections even at the $D$ scale. Both ratios, however, converge to 1 in the heavy-quark limit which gives us confidence that we control the heavy-quark-mass dependence of the various form factors. Furthermore, these ratios can help constrain the possible $q^2$ dependences of the various form factors around $q^2_{max}$ at $M{=}m_B$. \section{$\bar B^0\to\pi^+\ell^-\bar\nu$ and Dispersive Constraints} A second solution to the problem of the limited kinematical reach of lattice simulations of $heavy\to light$ quark decays is to combine lattice results for the relevant form factors around $q^2_{max}$ with dispersive bound techniques to obtain improved, model-independent bounds for the form factors for all $q^2$~\cite{btopi}. For the case of $\bar B^0\to\pi^+\ell^-\bar\nu$ decays, whose hadronic matrix element, $\langle \pi^+(p')|V^\mu|\bar B^0(p)\rangle$, is traditionally decomposed in terms of two form factors $f^+(q^2)$ and $f^0(q^2)$, one can use the kinematical constraint, $f^+(0)${=}$f^0(0)$, to further constrain the bounds. \subsection{Dispersive Bounds} The subject of dispersive bounds in semileptonic decays has a long history going back to S. Okubo {\it et al.}~who applied them to semileptonic $K\to\pi$ decays~\cite{SOkS71}. C. Bourrely {\it et al.}~first combined these techniques with QCD and applied them to semileptonic $D\to K$ decays \cite{CBoMR81}. Very recently, C.G. Boyd {\it et al.}~applied them to $B\to\pi\ell\bar\nu$ decays \cite{CBoGL95}. The starting point for $B\to\pi\ell\bar\nu$ decays is the polarization function \begin{eqnarray} \Pi^{\mu\nu}(q){=}i\int d^4x\ e^{iq\cdot x} \langle 0|T\left(V^\mu(x) V^{\nu\dagger}(0)\right)|0\rangle\nonumber\\ {=}(q^\mu q^\nu-g^{\mu\nu}q^2)\,\Pi_T(q^2)+q^\mu q^\nu \,\Pi_L(q^2) \ ,\label{twopoint} \end{eqnarray} where $\Pi_{T(L)}$ corresponds to the propagation of a $J^P{=}1^-\,(0^+)$ particle. The corresponding spectral functions, $\mbox{Im}\,\Pi_{T,L}$, are sums of positive contributions coming from intermediate $B^*$ ($J^P{=}1^-$), $B\pi$ ($J^P{=}0^+\mbox{ and }1^-$), $\ldots$ states and are thus upper bounds on the $B\pi$ contributions. Combining, for instance, the bound from $\mbox{Im}\,\Pi_L$ with the dispersion relation ($Q^2{=}-q^2$) \begin{eqnarray} \chi_L(Q^2)&=&\frac{\partial}{\partial Q^2} (Q^2\Pi_L(Q^2))\nonumber\\ &=&\frac{1}{\pi}\int_0^\infty dt\frac{t\,\mbox{Im}\,\Pi_L(t)}{\left(t+Q^2\right)^2} \ ,\label{eq:disprelS} \end{eqnarray} one finds \begin{eqnarray} \chi_L(Q^2)\ge\frac{1}{\pi}\int_{t^+}^\infty dt\,k(t,Q^2)|f^0(t)|^2 \ ,\label{eq:chilbnd} \end{eqnarray} where $t_{\pm}{=}(m_B\pm m_\pi)^2$ and $k(t,Q^2)$ is a kinematical factor. Now, since $\chi_L(Q^2)$ can be calculated analytically in QCD for $Q^2$ far enough below the resonance region (i.e. $-Q^2\ll m_b^2$), \eq{eq:chilbnd} gives an upper bound on the weighted integral of the magnitude squared of the form factor $f^0$ along the $B\pi$ cut. To translate this bound into a bound on $f^0$ in the region of physical $B\to\pi\ell\bar\nu$ decays is a problem in complex analysis (please see \re{btopi} for details). A similar constraint can be obtained from $\Pi_T$ for $f^+$. There, however, one has to confront the additional difficulty that $f^+$ is not analytic below the $B\pi$ threshold because of the $B^*$ pole. The beauty of the methods of \re{CBoMR81} is that they enable one to incorporate information about the form factors, such as their values at various kinematical points, to constrain the bounds. For the case at hand, however, these methods must be generalized in two non-trivial ways. In constructing these generalizations, one must keep in mind that the bounds: 1) form inseparable pairs; 2) do not indicate the probability that the form factor will take on any particular value within them. \subsection{Imposing the Kinematical Constraint} The first problem is that \eq{eq:chilbnd} and the equivalent constraint for $f^+$ yield independent bounds on the form factors which do not satisfy the kinematical constraint $f^+(0){=}f^0(0)$. The bounds on $f^+$ require $f^+(0)$ to lie within an interval of values $I_+$ and those on $f^0$, within an interval $I_0$. Together with these bounds, however, the kinematical constraint requires $f^+(0){=}f^0(0)$ to lie somewhere within $I_+\cap I_0$. Thus, we seek bounds on the form factors which are consistent with this new constraint. A natural definition is to require these new bounds to be the envelope of the set of pairs of bounds obtained by allowing $f^+(0)$ and $f^0(0)$ to take all possible values within the interval $I_+\cap I_0$. In \re{btopi}, I show how this envelope can be constructed efficiently and that the additional constraint can only improve the bounds on the form factors for all $q^2$. Also, as a by product, one obtains a formalism which enables one to constrain bounds on a form factor with the knowledge that it must lie within an interval of values at one or more values of $q^2$. \subsection{Taking Errors into Account} As they stand, the methods of \re{CBoMR81} can only accommodate exact values of the form factors at given kinematical points and contain no provisions for taking errors on these values into account. Of course, the results given by the lattice do carry error bars. More precisely, the lattice provides a probability distribution for the value of the form factors at various kinematical points. What must be done, then, is to translate this distribution into some sort of probability statement on the bounds. The conservative solution is to consider the probability that complete pairs of bounds lie within a given finite interval at each value of $q^2$. Then, using this new probability, one can define upper and lower $p\%$ bounds at each $q^2$ as the upper and lower boundaries of the interval that contains the central $p\%$ of this probability.\footnote{The density of pairs of bounds increases toward the center of the ``distribution'' as long as the distribution of the lattice results does.} These bounds indicate that there is at least a $p\%$ probability that the form factors lie within them at each $q^2$. \subsection{Lattice-Constrained Bounds} To constrain the bounds on $f^+$ and $f^0$, I use the lattice results of the UKQCD Collaboration \cite{DBuetal95}, to which I add a large range of systematic errors to ensure that the bounds obtained are conservative. Because of these systematic errors, the probability distribution of the lattice results is not known. I make the simplifying and rather conservative assumption that the results are uncorrelated and gaussian distributed. I construct the required probability by generating 4000 pairs of bounds from a Monte-Carlo on the distribution of the lattice results. My results for the bounds on the form factors are shown in \fig{fig:lcsr}. I have plotted the two form factors back-to-back to show the effect of the kinematical constraint. Without it, the bounds on $f^+$ would be looser, especially around $q^2{=}0$, where phase space is large. Since $f^+$ determines the rate, the kinematical constraint and the bounds on $f^0$ are important. Also shown in \fig{fig:lcsr} is the LCSR result of \re{VBeBKR95} which has two components: for $q^2$ below $15\mbox{ GeV}^2$, the $q^2$ dependence of $f^+$ is determined directly from the sumrule; for larger $q^2$, pole dominance is assumed with a residue determined from the same correlator. Agreement with the bounds is excellent. In \re{btopi}, the bounds are compared with the predictions of more authors as well as with direct fits of various parametrizations to the lattice results. Though certain predictions are strongly disfavored, the lattice results and bounds will have to improve before a firm conclusion can be drawn as to the precise $q^2$ dependence of the form factors. \begin{figure}[tb] \setlength{\epsfxsize}{64mm}\epsfbox[80 300 470 545]{lcsr.ps} \unit=0.8\hsize \point 0.95 -0.07 {\large{$q^2(\mbox{GeV}^2)$}} \point 0.20 0.55 {\Large{$f^0(|q^2|)$}} \point 0.770 0.55 {\Large{$f^+(q^2)$}} \caption{$f^0(|q^2|)$ and $f^+(q^2)$ versus $q^2$. The data points are the lattice results of UKQCD\protect\cite{DBuetal95} with added systematic errors. The pairs of fine curves are, from the outermost to the innermost, the 95\%, 70\% and 30\% bounds. The shaded curve is the LCSR result of \protect\re{VBeBKR95}.} \label{fig:lcsr} \end{figure} The bounds on $f^+$ also enable one to constrain the $B^*B\pi$ coupling $g_{B^*B\pi}$ which determines the residue of the $B^*$ pole contribution to $f^+$. The constraints obtained are poor because $f^+$ is weakly bounded at large $q^2$. Fitting the lattice results for $f^0$ and $f^+$ to a parametrization which assumes $B^*$ pole dominance for $f^+$ and which is consistent with HQS and the kinematical constraint gives the more precise result $g_{B^{*+}B^o\pi^+}=28\pm 4$.\footnote{The result of this fit is entirely compatible with our bounds on $f^+$ and $f^0$.} However, because this result is model-dependent, it should be taken with care. \subsection{Bounds on the rate and $|V_{ub}|$} As was done for the form factors, one can define the probability of finding a complete pair of bounds on the rate in a given interval and from that probability determine confidence level (CL) intervals for the rate. The resulting bounds are summarized in \tab{tab:rate}. They were obtained by appropriately integrating the 4000 bounds generated for $f^+(q^2)$, taking the skewness of the resulting ``distribution'' of bounds on the rate into account. \begin{table}[tb] \caption{Bounds on rate in units of $|V_{ub}|^2\,ps^{-1}$ and on $f^+(0)$. \label{tab:rate}} \begin{tabular}{ccc} \hline $\Gamma\left(\bar B^0\to\pi^+\ell^-\bar\nu\right)$ & $f^+(0)$ & CL\\ \hline $2.4\to 28$ & $-0.26\to 0.92$ & 95\% \\ $2.8\to 24$ & $-0.18\to 0.85$ & 90\% \\ $3.6\to 17$ & $0.00\to 0.68$ & 70\% \\ $4.4\to 13$ & $0.10\to 0.57$ & 50\% \\ $4.8\to 10$ & $0.18\to 0.49$ & 30\% \\ \hline \end{tabular} \end{table} The CL bounds obtained can be used, in conjunction with the branching ratio measurement of CLEO \cite{cleo}, to determine $|V_{ub}|$. One finds \begin{equation} |V_{ub}|10^4\sqrt{\tau_{B^0}/1.56\,ps}=(34\div 49)\pm 8\pm 6 \ ,\label{eq:vub} \end{equation} where the range given in parentheses is that obtained from the 30\% CL bounds on the rate and represents the most probable range of values for $|V_{ub}|$. The first set of errors is obtained from the 70\% CL bounds and the second is obtained by combining all experimental uncertainties in quadrature and applying them to the average value of $|V_{ub}|$ given by the 30\% CL results. This determination of $|V_{ub}|$ has a theoretical error of approximately 37\%. Though non-negligible, this error is quite reasonable given that the bounds on the rate are completely model-independent and are obtained from lattice data which lie in a limited kinematical domain and include a conservative range of systematic errors. \section{Conclusion and Outlook} Because HQS applies to $heavy\to light$ quark decays in a rather limited way, it is not possible to determine the full $q^2$ dependence of the relevant form factors from the lattice alone. The flip side of the coin is that the model-independent information provided by lattice calculations about these decays, though limited, is still very important, because the relevant matrix elements are not anchored at zero recoil by HQS, up to small radiative and power corrections, as they are in $heavy\to heavy$ quark decays. I have presented two approaches by which the information provided by the lattice on exclusive semileptonic $b\to u$ decays can be used to extract $|V_{ub}|$. Both approaches will benefit from forthcoming, improved lattice results. The lattice-constrained bounds would also benefit enormously from an increase in the range of accessible $q^2$. Finally, the techniques developed in \re{btopi} to construct lattice-improved bounds for $B\to\pi\ell\bar\nu$ decays are in principle applicable to limited results obtained by non-lattice means and to other processes such as $B\to\rho\ell\bar\nu$ and $B\to K^*\gamma$ decays.
proofpile-arXiv_065-485
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} The galactic microquasar GRS1915+105 is one of the most intriguing objects in astrophysics today. This object shows hard X-ray, radio, and infrared (IR) flares with variability on timescales from minutes to months (Mirabel and Rodriguez, 1996). In addition, VLA monitoring has revealed collimated ejection events exhibiting superluminal motion interpreted as synchrotron emission from jets with relativistic bulk motions (Mirabel and Rodriguez, 1994). Recently, Sams {\it et al.} (1996) have reported observations of extended near-IR ($2.2 \mu$m) emission near the beginning of a hard X-ray flare (Sazonov and Sunyaev, 1995). While the position angle of the near-IR extensions is consistent with the radio jets, the nature of the emission is not well-determined. Detailed multi-wavelength studies of the flaring of GRS1915+105 imply the presence of other gas and/or dust in the region surrounding the compact object, in addition to the jets (Mirabel and Rodriguez, 1996). Thus, possible sources of the extended IR flux include reprocessing of the hard X-ray flare on an ejected gas or dust disk, a wind outflow, or the radio jets, or synchrotron emission from the jets themselves (Sams {\it et al.}, 1996). \section{Observations} We observed GRS1915+105 with the COB infrared imager on the Kitt Peak National Observatory\footnote{KPNO is operated by AURA Inc. under contract to the National Science Foundation.} 2.1-m telescope on October 16 and 17, 1995, roughly 3 months after Sams {\it et al.} (1996). We used the K-band ($2.2 \mu$m) filter, with a 0.2-arcsec/pixel plate scale. On October 16, we took four 30-second exposures through light cirrus, with seeing of $\sim 0.7$-arcsec full-width half-maximum (FWHM). On October 17, we again took four 30-second exposures, this time under photometric conditions, but with $\sim 1.1$-arcsec FWHM seeing. During each sequence of exposures, we moved the telescope around the position of GRS1915+105 in order to avoid defects in the array. We subtracted the sky background and flatfielded each exposure, and then combined each set of four exposures. Figure 1 shows a contour map of the resulting image for the October 16 data. Photometry of the combined images gives K-band magnitudes of $K=13.51 \pm 0.05$ on October 16 and $K=13.43 \pm 0.03$ on October 17 (see also Eikenberry and Fazio, 1995). \section{Analysis} \subsection{Limits on point-like emission} In our analysis, we first concentrate on possible emission from a feature similar to that seen by Sams {\it et al.} (1996). While the separation between the stellar counterpart of GRS1915+105 and the extension of Sams {\it et al.} (1996) is less than 1/2 of our image FWHM, if the feature's brightness ($K=13.9$) remained constant it would be the source of 60\% of the photons in our images, and simple PSF-subtraction should reveal its presence. From each of the background-subtracted and flatfielded (but uncombined) 30-second exposures, we extract a $5 \times 5$-arcsec region centered on GRS1915+105. For our PSF, we extract an identical region centered on a star near GRS1915+105 with a very similar K-band flux (Star A in Figure 1). For each exposure, we scale the PSF and subtract it from the images of GRS1915+105. We see no evidence for extended structure in any of the PSF-subtracted exposures. We then combine the four PSF-subtracted exposures for each night. Again, we find no evidence for extended structure in the combined PSF-subtracted images (see Figure 2). We then apply a simple sliding-cell source-detection algorithm to the PSF-subtracted images. In this approach, we take a model PSF (a 2-D gaussian fit to Star A) and center it on a pixel in the image. We multiply the image pixel values by the corresponding PSF values at their location, and sum to obtain the total of the products. We then perform the same operation on the error image for the PSF-subtracted image (including the uncertainties in the PSF), summing the error products in quadrature. The ratio of the image sum to the error sum then gives the statistical significance of any point source at the image pixel location. Applying this algorithm to the PSF-subtracted images from both nights, we find no source with a statistical significance $>1 \sigma$ at any location. In order to estimate the sensitivity of our observations and analysis techniques, we perform a simple Monte Carlo simulation. First, we model the GRS1915+105 region as 2 point sources separated by 0.3-arcsec - the star plus a southern jet - as seen by Sams {\it et al.} (1996). We ignore the northern jet due to its much lower flux. Second, (using the model PSF for both point sources), we select a relative normalization of the model PSFs for the two point sources, scale and add them with the appropriate positional offsets, and then rescale the sum to give the same number of counts as in the real GRS1915+105 image. Next, we add normally-distributed random numbers (having standard deviations determined from the quadrature-summed uncertainties of both the GRS1915+105 and Star A (PSF) images) to each pixel of the simulated image. Finally, we take the image of Star A, scale it, and subtract it from the simulated image, exactly as with the actual GRS1915+105 images. In Figure 3, we present a typical simulated result of the PSF subtraction for an extended component with $K=13.9$. If the extended emission seen by Sams {\it et al.} (1996) had remained unchanged, we would have unambiguously found it in our data on both nights. In order to place an upper limit on any point-like flux at this position, we then decrease the flux of the extended emission in the model and repeat the simulation process. We set our upper limit on the flux of the extended component at the point where, for 100 Monte Carlo simulations as described above, we detect the extended component in the PSF-subtracted image at the 95\% confidence level using the sliding-cell source-detection algorithm. For October 16, the limit is $K>16.4$, while for October 17 (when the seeing was poorer), the limit is $K>15.8$. \subsection{Limits on extended emission} The limit on point-like emission is useful in confirming that the Sams {\it et al.} (1996) feature was transient, as expected for emission from the radio-emitting jets. However, since this feature is associated with the radio-emitting jets of GRS1915+105, it will exhibit superluminal motion, and may also expand at high velocities. Thus, we have performed further analyses searching for possible non-point-like emission from the jet. We perform this search using, once again, the sliding-cell algorithm described above. However, instead of using Star A as a PSF, we use a broadened PSF for the jet emissions. Applying the algorithm to the stellar-PSF-subtracted images, we find no evidence of extended emission using trial jet-PSFs with FWHM values of 0.8, 1.0, 1.25, and 1.5 arcsec. Given that the Sams {\it et al.} feature was point-like with their $\sim 0.2$ arcsec resolution, even if the feature expanded at $0.5 c$ (much faster than the limt for the radio-emitting jets), it could have expanded only to a FWHM of 1.25 arcsec, given the 12.5 kpc distance to GRS1915+105 (Rodriguez {\it et al.}, 1995). Thus, we conclude that their is no evidence for infrared jet emissions in our data. As with the point-like emission, we perform a Monte Carlo simulation to estimate the sensitivity of our observations and analysis techniques. We now assume that the jet has moved 1.2 arcsec farther from GRS1915+105 - an apparent velocity of $1.0c$, which is less than is observed in the radio jets (Mirabel and Rodriguez, 1994). We also assume a worst-case jet FWHM of 1.5 arcsec for the source-detection algorithm - an expansion velocity $>0.5c$, which is much greater than observed in the radio jets. This approach gives an upper limit of $K>17.7$ at the 95\% confidence level for any infrared jet emission. \section{Discussion} Given that the hard X-ray flaring activity which began in late June or early July 1995 (Sazonov and Sunyaev, 1995) continued through the time of these observations (Harmon {\it et al.}, 1995), the drop of a factor $>10$ in the $2.2 \mu$m flux at the observed location of Sams {\it et al.} (1996) places strong constraints on several of the proposed explanations for the extended near-IR emission. In particular, hypotheses involving the reprocessing of the X-ray emission on stellar winds, ejected dust or gas disks, or other steady-state or slow-moving structures do not appear to explain such behavior. This, in addition to the appearance of the IR features oppositely oriented about GRS1915+105, their position angle match with the radio jets, and the similarities of the North/South flux asymmetry to that in the radio (Sams et al., 1996), seems to confirm the identification of the features seen by Sams {\it et al.} as infrared jets. If the features are indeed due to infrared jets, then by the time of our observations, the southern (bright) jet would have moved more than 1 arcsec from GRS1915+105, and we have an upper limit of $K>17.7$ in this region. If the near-IR flux arises from reprocessing of the X-ray emission on the jet, the reprocessing efficiency may have dropped by this factor $>33$ due to the increased distance between the X-ray source and the jet and/or changes in the X-ray opacity of the jets. Alternatively, if the IR flux arises from synchrotron processes in the jet, then we can place an upper limit on the radiative lifetime of the IR-emitting particles, using the time separation between our observations and those of Sams {\it et al.} (1996) and the fact that our upper limit is a factor 33 lower in flux than the Sams {\it et al.} feature. Thus, we find that the $1/e$ radiative lifetime of the IR-emitting electroncs is $\tau <26$ days. For synchrotron emission, the relativistic electrons producing the IR emission have shorter radiative lifetimes than the radio-emitting electrons by a factor of $\sqrt{\nu_{IR} / \nu_{radio}} \sim 10^2$, independent of the magnetic field strength. Since the radio-emitting jets have lifetimes significantly less than 1 year, we find that our limits are compatible with the hypothesis that the Sams {\it et al.} feature arises from synchrotron processes in the radio-emitting jets. \section{Conclusions} We have presented near-infrared K ($2.2 \mu$m) band observations of the galactic microquasar GRS1915+105 on October 16 and 17, 1995 with a 0.2 arcsec/pixel plate scale under good seeing conditions. Using PSF subtraction of the stellar image of GRS1915+105, we find no evidence of near-infrared emissions as seen by Sams {\it et al.} (1996) in July, 1995. Simple modelling shows that we would have detected any such extended emission at the 95\% confidence level down to a limit of $K>16.4$, as compared to the $K=13.9$ jet observed by Sams {\it et al.} (1996). The fact that the IR flux at this location dropped by a factor $>10$ during a time when the hard X-ray flux increased seems to rule out reprocessing of the hard X-ray emission on slow-moving or steady-state structures near the compact object as a viable explanation for the extended IR emission, and confirms the hypothesis that the extended IR emission arises from the radio-emitting jets. If the features are indeed due to infrared jets, then by the time of our observations, the southern (bright) jet would have moved more than 1 arcsec from GRS1915+105, and we have an upper limit of $K>17.7$ in this region, a factor of $>33$ drop in the IR flux. This allows us to place an upper limit on the radiative lifetime of the feature of $\tau <26$ days. These limits are consistent with the hypothesis that the Sams {\it et al.} feature was due to synchrotron processes in the radio-emitting jets of GRS1915+105. \acknowledgements We would like to thank I.F. Mirabel for bringing the near-IR jets in GRS1915+105 to our attention, M. Merrill for assisting with the COB observations, and the anonymous referee for his/her helpful comments. S. Eikenberry is supported by a NASA Graduate Student Researchers Program fellowship through Ames Research Center.
proofpile-arXiv_065-486
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} This is the fourth paper of a series where the results of the EFAR project are presented. In Wegner et al. (1996, hereafter Paper I) the galaxy and cluster sample was described, together with the related selection functions. Wegner et al. (1997, hereafter Paper II) reports the analysis of the spectroscopic data. Saglia et al. (1997, hereafter Paper III) derives the photometric parameters of the galaxies. In this paper we describe the fitting technique used to derive these last quantities. A large number of papers have been dedicated to galaxy photometry. The reader should refer to the {\it Third Reference Catalogue of Bright Galaxies} (RC3, de Vaucouleurs et al. 1991) for a complete review of the subject. By way of introduction we give here only a short summary of the methods and tests adopted and performed in the past to derive the photometric parameters of galaxies. Using photoelectric measurements, photometric parameters have been derived by fitting curves of growth. The RC3 values are computed by choosing the optimal curve between a set of 15 (for $T=-5$ to $T=10$, see Buta et al. 1995), one for each type $T$ of galaxies. Photoelectric data are practically free from sky subtraction errors ($<0.5$\%), but can suffer from contamination by foreground objects. Typically, 5-10 data points are available per galaxy, with apertures which do not exceed 100 arcsec and do not always bracket the half-luminosity diameter. Burstein et al. (1987) (who fit the $R^{1/4}$ curve of growth to derive the photometric parameters of a set of ellipticals) discuss the systematic effects associated with these procedures. The total magnitudes $M_{TOT}$ and effective radii $R_e$ derived are biased depending on the set of data fitted. The errors in both quantities are strongly correlated, so that $\Delta \log R_e -0.3 \Delta \langle SB_e\rangle\approx$ constant, where $\langle SB_e \rangle = M_{TOT}+5\log R_e+2.5\log (2\pi)$ is the average surface brightness inside $R_e$. This constraint (Michard 1979, Kormendy \& Djorgovski 1989 and references therein) stems from the fact that the product $R{\langle I \rangle}^{0.8}$ varies only by $\pm$5\% for {\it all} reasonable growth curves (from $R^{1/4}$ to exponential laws) in a radius range $0.5R_e\le R\le 1.5 R_e$ (see Figure 1 of Saglia, Bender \& Dressler 1993). Here $\langle I \rangle$ is the average surface brightness inside $R$. If the galaxies considered are large ($R_e>10$ arcsec), no seeing corrections are needed (see Saglia et al. 1993). Until the use of CCD detectors, differential luminosity profiles of galaxies were obtained largely from photographic plates. The procedure required to calibrate the nonlinear response of the plates and to digitize them is very involved. As a consequence, it was possible to derive accurate luminosity profiles or two-dimensional photometry only for a small number of galaxies (see, for example, de Vaucouleurs \& Capaccioli 1979). Using this sort of data, Thomsen \& Frandsen (1983) derived $R_e$ and $M_{TOT}$ for a set of brightest elliptical galaxies in clusters at redshifts $<0.15$. They fit a two-dimensional $R^{1/4}$ law convolved with the appropriate point-spread function and briefly investigated the systematic effects of sampling (pixel size), signal-to-noise ratio, and shape of the profile on the derived photometric quantities. Lauberts \& Valentijn (1989) digitized and calibrated the blue and red plates of the ESO Quick Schmidt survey to derive the photometric parameters of a large set of southern galaxies. Here the total magnitudes are not corrected for extrapolation to infinity, but are defined as the integrated magnitude at the faintest measured surface brightness (beyond the 25 B mag arcsec$^{-2}$ isophote) for which the luminosity profile is monotonically decreasing. In addition, the catalogue gives the parameters derived by fitting a ``generalized de Vaucouleurs law'' ($I=I_0\exp(-(r/\alpha)^N)$; compare to Eq. \ref{r1n}) to the surface brightness profiles. The last 15 years have seen the increased use of CCDs for photometry. CCDs are linear over a large dynamic range, can be flatfielded to better than 1\% and allow one to eliminate possible foreground objects during the analysis of the data. Large samples of CCD luminosity profiles for early-type galaxies have been collected by Djorgovski (1985), Lauer (1985), Bender, D\"obereiner \& M\"ollenhoff (1988), Peletier et al. (1990), Lucey et al. (1991), J\o rgensen, Franx \& Kj\ae rgaard (1995). Using CCDs one can derive photometric parameters by fitting a curve-of-growth to the integrated surface brightness profile. One major concern of CCD photometry is sky subtraction. If the CCD field is not large enough compared to the half-luminosity radii of the galaxies, then the sky value determined from the frame may be systematically overestimated (due to contamination of the sky regions by galaxy light), leading to systematically underestimated $R_e$ and $M_{TOT}$. This problem might however be solved with the construction of very large chips or mosaics of CCDs (see MacGillivray et al. 1993, Metzger, Luppino \& Miyazaki 1995). Among the most recent studies of galaxy photometry is the Medium Deep Survey performed with the Hubble Space Telescope. Casertano et al. (1995) analyse 112 random fields observed with the Hubble Space Telescope Wide Field Camera prior to refurbishment to study the properties of faint galaxies. They construct an algorithm which fits the two-dimensional matrix of data points to perform a disk/bulge classification. The $R^{1/4}$ and exponential components are convolved with the point spread function (psf) of the HST and Monte Carlo simulations are performed to test the results. Disk-bulge decomposition is attempted only for a few cases (see Windhorst et al. 1994), because the data are in general limited by the relatively low signal-to-noise and by the spatial resolution. In order to derive total magnitudes, galaxy photometry involves extrapolation of curves of growth to infinity, and therefore relies on fits to the galaxy luminosity profiles. Recently, Caon, Capaccioli \& D'Onofrio (1993, hereafter CCO) and D'Onofrio, Capaccioli \& Caon (1994) focused on the use of the $R^{1/4}$ law to fit the photometry of ellipticals. CCO find a correlation with the galaxy size and argue that if an $R^{1/n}$ law (Sersic 1968, see Eq. \ref{r1n}) is used to fit the luminosity profiles, then smaller galaxies ($\log R_e[{\rm kpc}]<0.5$) are best fitted with exponents $1<n<4$, while larger ones ($\log R_e[{\rm kpc}]>0.5$) have $n>4$. Half-light radii and total magnitudes derived using these results may differ strongly from those using $R^{1/4}$ extrapolations. Finally, Graham et al. (1996) find that the extended shallow luminosity profiles of BCG are best fit by $R^{1/n}$ profiles with $n>4$. To summarize, the EFAR collaboration has collected photoelectric and CCD photometry for 736 galaxies (see Colless et al. 1993 and Paper III), 31\% of which appear to be spirals or barred objects. The remaining 69\% can be subdivided in cD-like (8\%), pure E (12 \%) and mixed E/S0 (49\%); the precise meaning of these classifications is explained in detail in \S 3.4 of Paper III. We derived circularly averaged luminosity profiles for all of the objects. Isophote shape analysis can only be reliably performed for the subset of our objects which are large and bright enough, and will be discussed in a future paper. Since 96\% of the EFAR galaxies have ellipticities smaller than 0.4, the use of circularized profiles does not introduce systematic errors on the photometric parameters derived (see \S \ref{decomposition}) and has the advantage of giving robust results for even the smaller, fainter objects in our sample. The galaxies show a large variety of profile shapes. Typically, each object has been observed several times, using a range of telescopes, CCD detectors, and exposure times, under different atmospheric and seeing conditions, with different sky surface brightnesses. Deriving homogeneous photometric parameters from the large EFAR data set has required the construction of a sophisticated algorithm to (i) optimally combine the multiple photoelectric and CCD data of each object, (ii) fit the resulting luminosity profiles with a model flexible enough to describe the observed variety of profiles, (iii) classify the galaxies morphologically and (iv) produce reliable magnitudes and half-luminosity radii. This paper describes our method as applied in Paper III. It explores the sources of random and systematic errors by means of Monte Carlo simulations, and develops a scheme to quantify the precision of the derived parameters objectively. The fitting algorithm searches for the best combination of the seeing-convolved, sky-corrected $R^{1/4}$ and exponential laws. This approach fulfils the requirement (ii) above: it produces convenient fits to the extended components of cD luminosity profiles, it models the profile range observed in E/S0 galaxies (from galaxies with flat cores to clearly disk-dominated S0s), and it reproduces the surface brightness profiles of spirals. Moreover, for the E/S0s and spirals, this approach determines the parameters of their bulge and disk components, to assist classification (requirement (iii)). Finally, this approach minimizes extrapolation (requirement (iv)) which is the main source of uncertainty involved in the determination of magnitudes and half-luminosity radii. Would it be possible to reach the same goals with another choice of fitting functions? We demonstrate here (\S \ref{profiles}) that the $R^{1/n}$ profiles quoted above can be seen as a ``subset'' of the $R^{1/4}$ plus exponential models and therefore might not meet requirement (ii). In addition, for $n>4$ they require large extrapolations and therefore might fail to meet requirement (iv). What is the physical interpretation of the two components of our fitting function? There are cases (the above cited cD galaxies and the galaxies with cores) where our two-component approach provides a good fitting function, but the ``disk-bulge'' decomposition is not physical. However, we argue that the systematic deviations from a simple $R^{1/4}$ law observed in the luminosity profiles of our early-type galaxies are the signature of a disk. We will investigate this question further in a future paper, where the isophote shape analysis of the largest and brightest galaxies in the sample will be presented. Would it be worth improving the present scheme by, for example, allowing for a {\it third} component (a second $R^{1/4}$ or exponential) to be fit? This could produce better fits to barred galaxies or to galaxies with cores and extended shallow profiles. However, it is not clear that the systematic errors related to extrapolation and sky subtraction could be reduced. Summarizing, the solution adopted here fulfils our requirements (i)-(iv). This paper is organized as follows. \S \ref{fitting} describes the three-step fitting technique. This involves the algorithm for the combination of multiple profiles of the same object (\S \ref{combination}), our two-component fitting technique with the additional option of sky fitting (\S \ref{diskbulge}), and the objective quality assessment of the derived parameters (\S \ref{quality}). \S \ref{montecarlo} presents the results of the Monte Carlo simulations performed to test the fitting procedure and assess the precision of the derived photometric parameters. We explore a large region of the parameter space ($R_{eB}, h$, D/B, $\Gamma$, see \S \ref{diskbulge} for a definition of the parameters) and test the performance of the fitting algorithm (\S \ref{parameter}). In \S \ref{sky} we investigate the systematic effects introduced by possible errors on sky subtraction and test the algorithm to correct for this effect (see \S \ref{diskbulge}). The influence of the limited radial extent of the profiles (\S \ref{extension}), of the signal-to-noise ratio (\S \ref{snratio}), and of seeing and pixellation (\S \ref{seeing}) are also investigated. The profile combination algorithm is tested in \S \ref{testcombination}. In \S \ref{decomposition} we assess the effectiveness of using the fitting algorithm to derive the parameters of bulge and disk components of a simulated galaxy. A number of different profiles are considered in \S \ref{profiles} to test their systematic effect on the photometric parameters. We show that the $R^{1/n}$ profiles can be reproduced by a sequence of $R^{1/4}$ plus exponential profiles, with small systematic differences ($<0.2$ mag arcsec$^{-2}$) over the radial range $R_e/20<R<5R_e$ (see discussion above). In \S \ref{discussion} we discuss how to estimate the precision of the derived photometric parameters. In \S \ref{conclusions} we summarize our results in terms of the expected uncertainties on the derived photometric parameters. \section{The fitting procedure} \label{fitting} The algorithm devised to fit the luminosity profiles of EFAR galaxies (see Paper III) involves three connected steps, (i) the combination of multiple profiles, (ii) the two-component fitting, and (iii) the quality estimate of the results. In the first step, the multiple CCD luminosity profiles available for each object are combined taking into account differences in sensitivity or exposure time, and sky subtraction errors. A set of multiplicative and additive constants is determined ($k_i$, $\Delta_i$), which describe respectively the relative scaling due to sensitivity and exposure time and the relative difference in sky subtraction errors. The absolute value of the scaling is the absolute photometric calibration of the images. This is accomplished as described in Paper III, making use of the photoelectric aperture magnitudes and absolute CCD calibrations. The absolute value of the sky correction can be fixed either to zero or to a percentage of the mean sky, or passed to the second step to be determined as a result of the fitting scheme. The second step fits these combined profiles. The backbone of the fitting algorithm is the sum of the seeing-convolved $R^{1/4}$ and the exponential laws. We have discussed the advantages of this choice in the Introduction. This combination produces a variety of luminosity profiles which can fit a large number of realistic profiles to high accuracy. The photometric parameters derived from this approach do not require large extrapolations, if the available profiles extend to at least $4R_e$. When galaxies with disk and bulge components (E/S0s and spirals) are seen at moderate inclination angles (as it is the case for the EFAR sample, where 96\% of galaxies have ellipticities less than 0.4, see Paper III), then the algorithm is also able, to some extent, to determine the parameters of the two components. In Paper III this information is used, together with the visual inspection of the images and, sometimes, the spectroscopic data, to classify each EFAR object as E, E/S0, or spiral. While we believe that in these cases the two components of the fits are indicative of the presence of two physical components, additional investigation is certainly required to test this conclusion. This will involve the isophote shapes analysis (Scorza \& Bender 1995), the fitting of the two-dimensional photometry (Byun \& Freeman 1995, de Jong 1996), the colors and metallicities (Bender \& Paquet 1995), and the kinematics (Bender, Saglia \& Gerhard 1994) of the objects. We intend to address some of these issues in future papers for a selection of large and bright EFAR galaxies. The third step assigns quality parameters to the derived photometric parameters. Several factors determine how accurate these parameters can be expected to be. \S \ref{montecarlo} explores in detail the effects of sky subtraction errors, radial extent, signal-to-noise ratio, seeing and sampling, and goodness of fit. A global quality parameter based on these results quantifies the precision of the final results. \subsection{Profile combination} \label{combination} The first step of the fitting algorithm is to combine the multiple profiles available for each galaxy. Fitting each profile separately, and averaging the results produces severely biased results if the fitted profiles differ in their signal-to-noise ratio, seeing and sampling, radial extent, and sky subtraction errors. Only a simultaneous fit can minimize the biasing effects of these factors (see \S \ref{testcombination}). Apart from the very central regions of galaxies, where seeing and pixel size effects can be important, the profiles of the same object taken with different telescopes and instruments differ by a normalization (or multiplicative constant) only and an additive constant. The first takes into account differences in the efficiency and transparency, while the second adjusts for the relative errors in the sky subtraction. Let $I_i(R)$, $i=1$ to $n$ denote the $n$ available profiles in counts per arcsec$^2$ at a distance $R$ from the center, and consider the profile $I_{max}(R)$ as the one having the maximum radial extent. In general the radial grids on which the profiles $I_i(R)$ have been measured will not be the same, but it will always be possible to (spline) interpolate the values of $I_{max}(R)$ on each of the grid points of the other profiles $I_i(R)$. The normalization $k_i$ of the profiles $I_i(R)$ relative to the profile $I_{max}(R)$ and the quantity $\Delta_i$ (related to $\Delta_i/k_i$ the correction to the sky value of the profile $I_i(R)$) are the multiplicative and additive constants to be sought, so that: \begin{equation} \label{corrected} I_i'(R)=k_iI_i(R)-\Delta_i. \end{equation} The $k_i$ and $\Delta_i$ constants are determined by minimizing the $\chi^2$-like functions (see the related discussion for Eq. \ref{chitot}): \begin{equation} \label{chid} \chi^2_i=\sum_{R>R_c} w_i(R)\left(I_{max}(R)-k_i I_i(R)+\Delta_i\right)^2. \end{equation} The inner cutoff radius $R_c$ is 6 arcsec or half of the maximum extent of the profile, if this is less than 6 arcsec. This cutoff minimizes the influence of seeing, while retaining a reasonable number of points in the sums. Here $w_i(R)=1/\sigma_i(R)^2$ are the relative weights of the data points, which are related to the expected errors for the profile $I_i$: \begin{equation} \label{weightw} \sigma_i(R)=\frac {\sqrt{G_i I_i(R)+G_i\hbox{Sky}_i+RON_i^2/S_i^2}} {\sqrt{2\pi R/S_i}}, \end{equation} where $S_i$, $G_i$ and $RON_i$ are the scale (in arcsec/pixel), the gain and the readout noise of the CCD used to obtain the profile $I_i$ (see Table 2 of Paper III). The denominator of Eq. \ref{weightw} assumes that all of the pixels in the annulus at radius $R\neq 0$ have been averaged to get $I(R)$ and therefore underestimates the errors if some pixels have been masked to delete background or foreground objects superimposed on the program galaxies. If $R=0$ (i.e., the central pixel) the following equation is used: \begin{equation} \label{weightw0} \sigma_i(R=0)={\sqrt{G_i I_i(R=0)+G_i\hbox{Sky}_i+RON_i^2/S_i^2}}. \end{equation} The weight in this fit monotonically increases with radius. The errors $\sigma_{\mu_i}$ on the surface brightness magnitudes $\mu_i=-2.5 \log I_i$ are related to Eqs. \ref{weightw} and \ref{weightw0} through: \begin{equation} \label{sigmai} \sigma_{\mu_i}=\frac{2.5 \sigma_i(R)\log e}{I_i(R)}, \end{equation} By requiring $\partial \chi^2_i/ \partial k_i=0$ and $\partial \chi^2_i/ \partial \Delta_i=0$ we solve the linear system for $k_i$ and $\Delta_i$. At this stage the relative sky corrections are known for all of the profiles except the most extended one. This last correction $\Delta_{max}$ can either be computed as part of the fitting program (see Eqs. \ref{chikmax} and \ref{chidmax}), or fixed to a given value. In \S \ref{montecarlo} the strategy of setting the mean percentage sky errors (for a given galaxy) to zero will be tested extensively against the above. For this case one requires: \begin{equation} \label{meanzero} \frac{\Delta_{max}}{\hbox{Sky}_{max}}+\sum_i \frac{\Delta_i}{k_i \hbox{Sky}_i}=0. \end{equation} In general, Eq. \ref{meanzero} is not a good choice and gives rise to systematic errors (see Fig. \ref{figskyerror}), however it is preferred when the sky fitting solution (Eq. \ref{chidmax}) requires excessively large extrapolations. Forty percent of the fits presented in Paper III are performed using Eq. \ref{meanzero}. Note that for both Eq. \ref{meanzero} and \ref{chidmax} described below, the value of $\Delta_{max}$ is determined iteratively, by minimizing Eq. \ref{chid}, having redefined $I_{max}(R)$ as $I'_{max}(R)$, where $I'_{max}(R)=I_{max}(R)-\Delta_{max}$, and repeating the procedure until it convergences. Four or five iterations are needed to reach a precision $<10^{-5}$ when Eq. \ref{meanzero} is used. Convergence is reached while performing the non-linear fitting of \S \ref{diskbulge}, when using Eq. \ref{chidmax}. Sky corrections, as computed in Paper III, are less than 1 \% for 80\% of the cases examined. The absolute scaling, $k_{max}$, of the $I_{max}(R)$ profile represents the photometric calibration of the profiles. This is performed as described in Paper III using the photoelectric aperture magnitudes and CCD zero-points. In the following we set $k_{max}=1$. \subsection{$R^{1/4}+$exponential law fitting} \label{diskbulge} The surface brightness profiles of each galaxy are modeled simultaneously by assuming that they can be represented by the sum of a de Vaucouleurs law (the ``bulge'' component indicated by B) and an exponential component (the ``disk'' component indicated by D): \begin{equation} \label{fittingfun} f(R,R_{eB},h,D/B,\Gamma,S)_{B+D}=f_B+f_D, \end{equation} where $R_{eB}$ is the half-luminosity radius of the bulge component, $h$ the exponential scale length of the disk component, $D/B$ the disk to bulge ratio, $\Gamma$ the FWHM of the seeing profile, and $S$ the pixel size. Both laws are seeing-convolved as described by Saglia et al. (1993) and take into account the effects of finite pixel size. Definitions and numerical details can be found in the Appendix. The results presented in Paper III show that Eq. \ref{fittingfun} gives fits with respectably small residuals. The differences in surface brightness $\Delta \mu=\mu-\mu_{fit}$ are typically less than 0.05 mag arcsec$^{-2}$, while those between the integrated aperture magnitudes are a factor two smaller. However our formal values of reduced $\chi^2$ (see discussion below) indicate that very few galaxies (less than 10\%) have luminosity profiles that are fit well by the model disk and bulge. Over 90\% of the fits have reduced $\chi^2$ larger than 2. In this sense Eq. \ref{fittingfun} is not a statistically good representation of the galaxy profiles. A hybrid non-linear least squares algorithm is used to find the $R_{eB}$, $h$, D/B and the vector of seeing values which gives the best representation $f_{B+D}(R)$ of the profiles $I_i(R)$, taking into account the sky corrections $\Delta_i/k_i$. The algorithm uses the Levenberg-Marquardt search (Press et al. 1986), repeated several times starting from randomly scattered initial values of the parameters. The search is repeated using the Simplex algorithm (Press et al. 1986). The best of the two solutions found is finally chosen. This approach minimizes the biasing influence of the possible presence of several nearly-equivalent minima of Eq. \ref{chitot}, a problem present especially when low disk-to-bulge ratios are considered (see discussion in \S \ref{parameter}). All of the profiles $I_i(R)$ available for a given galaxy are fitted simultaneously determining the appropriate value of the seeing $\Gamma_i$, for each single profile $i$. The minimization is performed on the function: \begin{equation} \label{chitot} \chi^2_{totB+D}=\sum_i\left(\sum_{R,\lambda_i f_{B+D}>-\Delta_i/k_i} T_>^2+\sum_{R,\lambda_i f_{B+D}<-\Delta_i/k_i} T_<^2\right), \end{equation} where: \begin{equation} \label{termmaj} T_>=-2.5\log \left[\frac{\lambda_i f_{B+D}(R,R_{eB},h,D/B,\Gamma_i,S_i)+\Delta_i/k_i}{I_i(R)}\right]\frac{p_i}{\sigma_{\mu_i}}, \end{equation} and: \begin{equation} \label{termmin} T_<=-2.5 \log \left[\frac{\lambda_i f_{B+D}(R,R_{eB},h,D/B,\Gamma_i,S_i)} {I_i(R)-\Delta_i/k_i}\right]\frac{p_i}{\sigma_{\mu_i}} \end{equation} The penalty function $p_i$ is introduced to avoid unphysical solutions and increases $\chi^2_{totB+D}$ to very large values when $D/B<0$ or when the values of of $R_{eB}$ or $h$ become too large ($>300''$) or too small ($<1''$). The use of the $T_>$ and $T_<$ terms ensures that the arguments of the logarithm are always positive. The sky correction is usually applied to the fitting function (see Eq. \ref{termmaj}). However, data points where $\lambda_i f_{B+D}+ \Delta_i/k_i<0$ (this may happen when a negative sky correction $\Delta_i/k_i$ is applied) are included using Eq. \ref{termmin}, which applies the sky correction to the data points. Note also that Eq. \ref{chitot} is the weighted sum of the squared {\it magnitude} residuals. This is to be preferred to the weighted sum of the squared linear residuals, which is dominated by the data points of the central parts of the galaxies. The model normalization relative to the profile $I_i(R)$, $\lambda_i$, is determined by requiring $\partial \chi^2_{\lambda_i}/ \partial \lambda_i=0$, where: \begin{equation} \label{chikmax} \chi^2_{\lambda_i}=\sum_R w_i(R)\left(I_i(R)-\lambda_i f_{B+D}(R)- \Delta_i/k_i\right)^2. \end{equation} Note that the ratios $\lambda_{max}/\lambda_i$ can in principle differ from the constants $k_i$, because of (residual) seeing effects (see, e.g., $R_c$ in Eq. \ref{chid}) and systematic differences between model and fitted profiles. In fact, the differences are smaller than 8\% in 85\% of the fits performed with more than one profile (see Paper III). When a bulge-only or a two-component model is used, the total magnitude of the fitted galaxy, in units of the $I_{max}(R)$ profile, is computed as $M_{TOT}=-2.5 \log (L_B+L_D)$, where $L_B=\lambda_{max} R_{eB}^2$ (see Eq. \ref{bulge}, with this normalization one has $I_{eB}=\lambda_{max}/(7.22\pi)$) is the luminosity of the bulge and $L_D=(D/B)L_B$ is the luminosity of the disk. When a disk-only model is used, then $M_{TOT}=-2.5 \log L_D$, where $L_D=\lambda_{max}h^2$ (see Eq. \ref{disk}, with this normalization one has $I_0=\lambda_{max}/(2\pi)$). Note again that the photometric calibration of these magnitudes $M_{TOT}$ to apparent magnitudes $m_{T}$ is performed in Paper III using photoelectric aperture magnitudes and CCD zero-points. The sky correction to the profile $I_{max}$ can be set to a given value (zero for no sky correction, using Eq. \ref{meanzero} for zero mean percentage sky correction). Alternately, a fitted sky correction $\Delta_{max}$ can be determined by additionally requiring $\partial \chi^2_{\lambda_{max}}/ \partial \Delta_{max}=0$, where: \begin{equation} \label{chidmax} \chi^2_{\lambda_{max}}=\sum_R w_{max}(R)\left(I_{max}(R)-\lambda_{max}f_{B+D}(R)- \Delta_{max}\right)^2. \end{equation} If the resulting $\Delta_{max}$ produces $\lambda_{max}f_{B+D}+\Delta_{max}<0$ at any $R$, Eq. \ref{termmin} is used to compute the corresponding contribution to Eq. \ref{chitot}. When using Eq. \ref{chidmax}, the constants $k_i$ and $\Delta_i$ are computed again using $I'_{max}(R)=I_{max}(R)-\Delta_{max}$ (see the profile combination iterative algorithm in \S \ref{combination}). The Monte Carlo simulations of \S \ref{montecarlo} show that Eq. \ref{chidmax} gives an unbiased estimate of the sky corrections when the $f_{B+D}$ is a good model of the fitted profiles. Eq. \ref{meanzero} is to be preferred when large extrapolations are obtained; 60\% of the fits presented in Paper III are performed using Eq. \ref{chidmax}. One might use the equivalent of Eq. \ref{chidmax} for the profiles $I_i(R)$ to compute the corrections $\Delta_i$ directly from the fit, without having to go through Eq. \ref{chid}. This would automatically take into account the seeing differences of the profiles. However, tests show that this approach does not produce the correct relative sky corrections between the profiles, if the fitting function does not describe the fitted profiles well. Finally, one might try deriving $\lambda_i$ and $\Delta_i$ by minimizing Eq. \ref{chitot} for these two additional parameters. The adopted solution, however, speeds up the CPU intensive, non-linear minimum search, since $\lambda_i$ and $\Delta_i$ are computed analytically. The fit is repeated using a pure de Vaucouleurs law (D/B=0) and a pure exponential law (B/D=0). In analogy with Eq. \ref{chitot}, two other $\chi^2_{tot}$ are considered for these fits, $\chi^2_{totB}$ and $\chi^2_{totD}$. A (conservative) $3\sigma$ significance test (see discussion after Eq. \ref{chido}) is performed to decide whether the addition of the second component improves the fit significantly. The bulge-only fit is taken if: \begin{equation} \label{chibo} \frac{\chi^2_{totB}}{\chi^2_{totB+D}}-1<3\sqrt{\frac{2}{N^{free}_{B+D}}}. \end{equation} The disk-only fit is taken if: \begin{equation} \label{chido} \frac{\chi^2_{totD}}{\chi^2_{totB+D}}-1<3\sqrt{\frac{2}{N^{free}_{B+D}}}. \end{equation} The number of degrees of freedom of the $R^{1/4}$ plus exponential law fit is $N^{free}_{B+D}=N_{data}-N_{sky}-3-2N_{prof}$, where $N_{data}$ is the number of data points involved in the sum of Eq. \ref{chitot}, $N_{sky}=1$ if the sky fitting is activated, zero otherwise, and $3+2N_{prof}$ are the number of parameters fitted ($R_{eB},h,D/B$, $M_{TOT}$, $N_{prof}$ seeing values and $N_{prof}-1$ normalization constants $\lambda_i$, where $N_{prof}$ is the number of fitted profiles). If the errors $\sigma_{\mu_i}$ are gaussian, $\chi^2_{totB+D}$ follows a $\chi^2$ distribution of $N^{free}_{B+D}$ degrees of freedom. If the bulge plus disk model is a good representation of the data, then the $\chi^2_{totB+D}\approx N^{free}_{B+D}$ in the mean, with an expected dispersion $\sqrt{2N^{free}_{B+D}}$. In this case Eqs. \ref{chibo} and \ref{chido} are a 3$\sigma$ significance test on the conservative side, meaning that one-component models are preferred, if two-component models do not improve the fit by more than $3\sigma$. In fact, Paper III shows that only 10\% of the fits are statistically ``good'' ($\chi^2_{totB+D}\approx N^{free}_{B+D}$). The median reduced $\chi^2$, $\hat\chi^2=\chi^2_{totB+D}/ N^{free}_{B+D}$, is $\approx 6$, indicating the existence of statistically significant systematic deviations from the simple two-component models of Eq. \ref{fittingfun}. Fortunately, the tests performed in \S \ref{montecarlo} show that reliable photometric parameters can be obtained even in these cases. Note that fits based on the $R^{1/n}$ profiles {\it do not} give better results: Graham et al. (1996) obtain reduced $\chi^2\approx 10$ for their sample of brightest cluster galaxies. Eqs. \ref{chibo} and Eq. \ref{chido} as applied in Paper III select a bulge-only fit in 14 \% of the cases, and a disk-only fit in less than 1\%. In the 85\% of the cases when both components are used, the median value of the significant test is 16$\sigma$, with significance larger than 5$\sigma$ in 90\% of the cases. In the following sections and plots we shall indicate the reduced $\chi^2$ with $\chi^2$. Total magnitudes of galaxies are extrapolated values. In order to quantify the effect of the extrapolation, we also derive the percentage contribution to $M_{TOT}=-2.5 \log (L_B+L_D)$ due to the extrapolated light beyond the radius $R_{max}$ of the last data point. In 80\% of the galaxies examined in Paper III this extrapolation is less than 10\%. The half-luminosity radius $R_e$ (and the $D_n$ diameter, see Paper III) of the best-fitting function is computed using Eqs. \ref{bulgegrowth} and \ref{diskgrowth}, so that seeing effects are taken into account. Finally, the contamination of the sky due to galaxy light is estimated by computing the mean surface brightness in the annulus with radii $R_i^{max}$ and $2R_i^{max}$, where $R_i^{max}$ is the radius of the last data point of the profile $i$. Galaxy light contamination is less than 0.5 \% of the sky in 80\% of the cases studied in Paper III. Using the appropriate seeing-convolved tables (see \S \ref{diskbulge}), the fitting algorithm can also be used to fit a $f_\infty$ $\Psi=12$ model (see description in Saglia et al. 1993 and the appendix here) plus exponential, or a smoothed $R^{1/4}$ law plus exponential. These additional fitting models are useful to study the effects of the central concentration and radial extent of galaxies (see \S \ref{seeing}). \subsection{Quality parameters} \label{quality} The third step in the fitting procedure assigns quality estimates to the derived photometric parameters. Several factors determine their expected accuracy. (i) Low signal-to-noise images provide fits with large random errors. (ii) Images of small galaxies observed under poor seeing conditions and/or with inadequate sampling (a detector with large pixel size) give systematically biased fits. (iii) Images of large galaxies taken with a small detector give profiles with too little radial extent and fits involving large, uncertain extrapolations. (iv) Sky subtraction errors bias the faint end of the luminosity profiles and therefore the fitted parameters. Finally, (v) bad fits to the luminosity profiles provide biased quantities. The effects of these possible sources of errors are estimated by means of Monte Carlo simulations in \S \ref{montecarlo}. Based on these results, one can assign the quality estimates $Q_{max}$, $Q_\Gamma$, $Q_{S/N}$, $Q_{\hbox{Sky}}$, $Q_{\delta \hbox{Sky}}$, $Q_{E}$, $Q_{\chi^2}$ according to the rules listed in Table \ref{tabquality}, where increasing values of the quality estimates correspond to decreasing expected precision of the photometric parameters derived from the fits. The global quality parameter $Q$: \begin{equation} \label{qtot} Q=Max(Q_{max},Q_{\Gamma},Q_{S/N},Q_{\hbox{Sky}},Q_{\delta \hbox{Sky}},Q_E,Q_{\chi^2}), \end{equation} assumes values $1,2,3$, corresponding to expected precisions on total magnitudes $\Delta M_{TOT} \approx 0.05,0.15,0.4$, on the logarithm of the half-luminosity radius $\Delta \log R_e \approx 0.04, 0.1,0.3$ and on the combined quantity $FP=\log R_e-0.3\langle SB_e \rangle$ $\Delta FP \approx 0.005, 0.01,0.03$ (see \S \ref{discussion}, Fig. \ref{figqualdis}). Paper III shows that 16\% of EFAR galaxies have $Q=1$, 73\% have $Q=2$ and 11\% $Q=3$. Note that $M_{TOT}$ and, therefore, $FP$ are subject to the additional uncertainty due to the photometric zero-point. In Paper III we extensively discuss this source of error and find that it is smaller than 0.03 mag per object, for all of the cases (86\%) where a photoelectric or a CCD calibration has been collected. \begin{table}[ph] \caption[*]{The definition of the quality parameters} \label{tabquality} \begin{tabular}{cccccccc} &&&&&&&\\ \tableline \tableline $R_{max}/R_e^f$ & $Q_{max}$ & $R_e^f/\Gamma^f$ & $Q_\Gamma$ & $S/N$ & $Q_{S/N}$ & Extrap & $Q_E$ \\ \tableline $\le 1$ & 3 & $\le 2$ & 2 & $\le 300$ & 2 & $\ge0.3$ & 3 \\ $>1$,$\le2$ & 2 & $>2$ & 1 & $>300$ & 1 & $<0.3$ & 1 \\ $>2$ & 1 & & & & & & \\ \hline &&&&&&&\\ \tableline \tableline $\chi^2$ & $Q_{\chi^2}$ & $\mu_{\hbox{Sky}}-\langle SB_e^f \rangle$ & $Q_{\hbox{Sky}}$ & $|\delta \hbox{Sky/Sky}|$ & $Q_{\delta \hbox{Sky}}$ & & \\ \tableline &&&&&&&\\ $\ge25$ & 3 & $\le0.75$ & 2 & $>0.03$ & 3 & &\\ $\ge12.5$, $<25$ & 2 & $>0.75$ & 1 & $>0.01$,$<0.03$ & 2 & &\\ $<12.5$ & 1 & & & $<0.01$ & 1 & &\\ \tableline \end{tabular} \end{table} \section{Monte Carlo Simulations} \label{montecarlo} The fitting procedure described in the previous section has been extensively tested on simulated profiles with the goals of checking the minimization algorithm and quantifying the effects of the errors described in \S \ref{quality}. Luminosity profiles of models with known parameters have been fitted, to compare input and output values. In all of the following figures the output parameters of the fit are indicated with the superscript $f$ for ``fit'' (for example, $\Gamma^f$). As a first step (\S \ref{parameter}-\ref{testcombination}, Figures \ref{figresdb}-\ref{figcombination}), we ignore possible systematic differences between test profiles and fitting functions (such as the ones possibly present when fitting real galaxies, see discussion in Paper III) and generate a number of $R^{1/4}$ plus exponential model profiles of specified $R_{eB}$, $h$, $D/B$ ratio, seeing $\Gamma$ and total magnitude, using the seeing-convolved tables described in the Appendix. A constant can be added (subtracted) to simulate an underestimated (overestimated) sky subtraction. Given the pixel size, the sky value, the gain and readout noise, appropriate gaussian noise is added to the model profile following Eqs. \ref{weightw} and \ref{weightw0}. The maximum extent of the profiles can be specified to simulate the finite size of the CCD. The profile is truncated at the radius where noise (or the sky subtraction error) generates negative counts for the model. The signal-to-noise ratios computed in the following refer to the total number of counts in the model profile out to this radius. The parameter space explored in all of the simulations discussed in \S \ref{parameter}-\ref{seeing} is displayed in Figure \ref{figparspace} and covers the region where the EFAR galaxies are expected to reside (see Paper III). Different symbols identify the models (see caption of Fig. \ref{figparspace}). As a second step (\S \ref{decomposition}-\ref{profiles}, Figures \ref{figbulgedisk}-\ref{figr1npar}), we explore the influence of systematic differences between test profiles and fitting functions. In \S \ref{decomposition} we show that fitting circularized profiles of moderately flattened galaxies (as the ones observed in Paper III) allows good determinations of the photometric parameters and also of the bulge and disk components. In \S \ref{profiles} we fit the $R^{1/n}$ profiles, achieving two results. First, we quantify the influence of the quoted systematic effects on the fitted photometric parameters. Second, we suggest that the possible correlation between galaxy sizes and exponent $n$ (see discussion in the Introduction) reflects the presence of a disk component in early-type galaxies. \S \ref{discussion} summarizes the results by calibrating the quality parameter $Q$ of Eq. \ref{qtot}. \subsection{The parameter space} \label{parameter} In this section we discuss the results obtained by fitting the models indicated by the crosses in Figure \ref{figparspace}. For clarity the parameters are also given in Table \ref{tabpara}. No sky subtraction errors are introduced and the sky correction algorithm is not used. The detailed analysis of the possible sources of systematic errors discussed in \S \ref{sky}-\ref{snratio} is performed on the same sample of models. More extreme values of the parameters are used when testing the effects of seeing and resolution (\S \ref{seeing}). The profiles tested in this section extend out to $4 R_e$, have a pixel size of 0.4 arcsec and normalization of $10^7$ counts, with $G_i=RON_i=1$ (see Eq. \ref{weightw}), corresponding to $S/N\approx 1000$. \begin{table}[hp] \centering \caption[*]{The parameters of the models indicated by the crosses of Figure \ref{figparspace} (see \S \ref{parameter}). A model for each combination of parameters in the two blocks separately has been generated. $D/B=\infty$ indicates exponential models ($B/D=0$). } \label{tabpara} \begin{tabular}{ll} &\\ \tableline \tableline Parameter & Values \\ \tableline &\\ $R_{eB}('')$ & 4, 8, 12, 16, 20, 32\\ $h('')$ & 4, 8, 12, 16, 20, 32\\ $D/B$ & 0, 0.1, ..., 1, 1.2, 1.6, 2, 3.2, 5, $\infty$\\ $\Gamma('')$ & 1.5, 2.5\\ Sky/pixel & 1000 \\ &\\ \tableline &\\ $R_{eB}('')$ & 2, 3\\ $h('')$ & 3, 6\\ $D/B$ & 0, 0.2, 0.4, 0.8, 1.6, $\infty$\\ $\Gamma('')$ & 1.5, 2.5\\ Sky/pixel & 500 \\ &\\ \tableline \end{tabular} \end{table} Figure \ref{figresdb} shows the precision of the reconstructed parameters. Total magnitudes are derived with a typical accuracy of 0.01 mag, $R_e$ and $\Gamma$ to 3\%, $R_{eB}$ and $h$ to $\approx 8$\%, $D/B$ to $\approx 10$\%. The errors $\Delta M_{TOT}=M_{TOT}-M_{TOT}^f$ and $\Delta R_e =\log R_e/R_e^f$ are highly correlated, with insignificant differences from the relation $\Delta FP = \Delta R_e-0.3(\Delta M_{TOT}+5\Delta R_e)= \Delta R_e-0.3\Delta \langle SB_e \rangle$. Galaxies with faint ($D/B<0.3$) and shallow ($h/R_{eB}>2$) disks show the largest deviations. This partly reflects a residual (minimal) inability of the fitting program to converge to the real minimum $\chi^2$ (there are 3 points with $\chi^2>10$), but stems also from the degeneracy of the Bulge plus Disk fitting. Figure \ref{figbadfit} (a) and (b) show the example of a model with $D/B=0.1$ and $h/R_{eB}=5$ where a very good fit is obtained ($\chi^2=1.3$) yet there is a 0.05 mag error on $M_{TOT}$ and the disk solution is significantly different from the input model. Note that the largest deviations $\Delta M_{TOT}$ and $\Delta R_e$ are associated with the largest extrapolations ($\approx 20$\%). In the following sections we shall see that extrapolation is the main source of uncertainty, when determining total magnitudes and half-luminosity radii. The uncertainties $\Delta R_{eB}$ on the bulge scale length are smallest with bright bulges, while those on the disk scale length $\Delta h$ are smallest with bright disks. The algorithm to opt for one-component best-fits (Eqs. \ref{chibo}-\ref{chido}) identifies successfully all of the one-component models tested (bulges plotted at $\log D/B =-1.1$ and $\log h/R_{eB}=-1.1$, disks plotted at $\log D/B =1.1$ and $\log h/R_{eB}=1.1$ in Figure \ref{figresdb}). For only two models (with $D/B=0.1$ and large $h/R_{eB}$) is the bulge-only fit preferred (using the 3$\sigma$ test) to the two-component fit (circled points in Figure \ref{figresdb}). Figure \ref{figquatexp} shows the results obtained by fitting a pure bulge or a pure disk. As before, no sky subtraction error is introduced and the sky correction algorithm is not activated. Neglecting one of the two components strongly biases the derived total magnitudes and half-luminosity radii. In the case of the $R^{1/4}$ fits, already test models with values of D/B as small as $\approx 0.2$ give fitted magnitudes wrong by 0.2 mag, and $R_e$ by more than 30\%. The systematic differences correlate with the amount of extrapolation involved, and large extrapolations yield strongly overestimated magnitudes and half-luminosity radii. However, the resulting correlated errors $|\Delta FP|$ are almost always smaller than 0.03. In the case of pure exponential fits, the derived total magnitudes and half-luminosity radii are always smaller than the true values, since very little extrapolation ($<1$\%) is involved. Consequently, a positive, correlated error $\Delta FP$ ($\approx +0.03$) is obtained. Finally, note that pure bulge fits are {\it bad} fits of the surface brightness profiles ($\chi^2>10$), but may appear to give acceptable fits of the integrated magnitude profiles. (One can easily show that the differences in integrated magnitudes are the weighted mean of the differences in surface brightness magnitudes). Figure \ref{figbadfit} (c) and (d) shows such an example for an $R^{1/4}$ fit to a model with $D/B=0.8$ and $h/R_{eB}=1$. The residuals in the integrated magnitude profile are always smaller than 0.07 mag, but a $\chi^2=181$ is derived, with $\Delta M_{TOT}=0.32$ and $R_e^f/R_e=1.65$. These considerations suggest that magnitudes and half-luminosity radii derived by fitting the $R^{1/4}$ curve of growth to integrated magnitude profiles (Burstein et al. 1987, Lucey et al. 1991, J\o rgensen et al. 1995, Graham 1996) may be subject to systematic biases, as indeed Burstein et al. (1987) warn in their Appendix. This might be important for the sample of J\o rgensen et al. (1995), where substantial disks are detected in a large fraction of the galaxies by means of an isophote shape analysis. It is certainly very important for the sample of cD galaxies studied by Graham (1996, see discussion in Paper III). These objects have luminosity profiles which differ strongly from an $R^{1/4}$ law. Finally, note that the {\it systematic} errors shown in Figure \ref{figquatexp} (and in the figures of the following sections) cannot be simply estimated by considering the shape of the $\chi^2$ function near the minimum. Figure \ref{figcontour} shows the $1,2,3,5\sigma$ contours of constant $\chi^2$ for an $R^{1/4}$ fit to a $h/R_{eB}=0.5$, $D/B=0.1$ $R^{1/4}$ plus exponential model. The reduced $\chi^2$ (8.47 at the minimum) has been normalized to 1, so that $1\sigma$ corresponds to a (normalized) $\chi^2=1+\sqrt{2/N_{free}}=1.11$. The errors, estimated at the $5\sigma$ contour, underestimate the differences between the fit and the model by a factor 2. This results from the extrapolation involved and can be as large as one order of magnitude for models with larger D/B ratios. \begin{figure} \plotone{f1.eps} \caption[f1.eps]{The parameter space of the $R^{1/4}$ plus exponential profile of the Monte Carlo simulations discussed in Figure \ref{figresdb}-\ref{figgamma}. Models of Figs. \ref{figresdb}-\ref{figskyfix}: crosses (see also Table 2). Models of Fig. \ref{figextension}: skeletal triangles. Models of Fig. \ref{figflux}: open triangles. Models of Figs. \ref{figseeing}-\ref{figgamma}: open squares. Models of Fig. \ref{figcombination}: open pentagons. Models of Fig. \ref{figbulgedisk}: open hexagons. The small dots show the position of the EFAR galaxies as determined in Paper 3. The parameters of bulge only models are shown with $h=0$. The parameters of disk only models are shown at $R_{eB}=0$. See discussion in \S \ref{montecarlo}.} \label{figparspace} \end{figure} \begin{figure} \plotone{f2.eps} \caption[f2.eps]{The reconstructed parameter space for the models indicated by the crosses in Fig. \ref{figparspace}. No sky error is present. The quantities plotted on the y-axis are defined as $\Delta M_{TOT}=M_{TOT}-M_{TOT}^f$, $\Delta R_e=\log R_e/R_e^f$, $\Delta FP = \Delta R_e-0.3(\Delta M_{TOT}+5\Delta R_e)=\Delta R_e-0.3\Delta \langle SB_e \rangle$, $\Delta (D/B)=\log [(D/B)/(D/B)^f$], $\Delta R_{eB}=\log R_{eB}/R_{eB}^f$, $\Delta h = \log h/h^f$, $\Delta \Gamma = \log \Gamma/\Gamma^f$. On the x-axis, the first three boxes show the input parameters of the models in the logarithm units ($\log D/B$, $\log h/R_{eB}$, $\log R_e/\Gamma$). The last three boxes show the differences in magnitudes between the assumed sky value and the average effective surface brightness of the models ($\mu_{\hbox{\rm Sky}}-\langle SB_e \rangle$), the logarithm of the reduced $\chi^2$, and the fraction of light extrapolated beyond $R_{max}$ used in the determination of $M^f_{TOT}$. Models with $D/B=0$ (pure $R^{1/4}$ laws) are plotted at $\log D/B=-1.1$ and $\log h/R_{eB}=-1.1$. Models with $B/D=0$ (pure exponential laws) are plotted at $\log D/B=1.1$ and $\log h/R_{eB}=1.1$. Models with $D/B\neq 0$ which have been fitted with one component are circled. See \S \ref{parameter} for a discussion of the results.} \label{figresdb} \end{figure} \begin{figure} \plotone{f3.eps} \caption[f3.eps]{(a) A circular disk plus bulge model with $D/B=0.1$ and $h/R_{eB}=5$ (crosses). The dotted curves show the luminosity profiles $\mu(R)=-2.5\log I(R)$ of the bulge and the disk components, the dashed curve the fitted disk component. (b) The differences $\Delta \mu$ (in mag arcsec$^{-2}$, open squares) between model surface brightness and the fitted one (dotted curve) (see \S \ref{parameter}). (c) The $R^{1/4}$ fit (solid curve) to the surface brightness magnitude profile of a circular disk plus bulge model with $D/B=0.8$ and $h/R_{eB}=1$ (crosses, one point in every four). (d) The differences $\Delta mag$ between the $R^{1/4}$ integrated magnitudes and the fitted ones (solid curve, see \S \ref{parameter}). Note that $|\Delta mag| <0.07$ even if large deviations $\Delta \mu$ are present. (e) The fit to the circularized profile of a flattened bulge plus an inclined disk model (see \S. \ref{decomposition}). The luminosity profile of the model (crosses, one point in every four; the bulge and the disk components, with the listed parameters, are the full curves) is best fitted by an $R^{1/4}$ plus exponential law (dashed curves) with parameters $R_{eB}^f=16.34$ arcsec, $h^f=13.93$ arcsec, $(D/B)^f=0.13$, $R_e=17.51$ arcsec. (f) The residuals $\Delta \mu$ of the fit (open squares, one point in every four) and the differences between the growth curves $\Delta mag$ (full curve).} \label{figbadfit} \end{figure} \begin{figure} \plotone{f4.eps} \caption[f4.eps]{The effects of fitting disk plus bulge test profiles by either a single bulge (first three rows of plots) or a single disk (last three rows of plots) model. The test models are indicated by the crosses of Fig. \ref{figparspace}. $\Delta M_{TOT}$, $\Delta R_e$, and $\Delta FP$ are defined as in Fig. \ref{figresdb}, x-axis as in Fig. \ref{figresdb}. See \S \ref{parameter} for the a discussion of the results. Note the change of scale on the ordinate axis with respect to Figure \ref{figresdb}.} \label{figquatexp} \end{figure} \begin{figure} \plotone{f5.eps} \caption[f5.eps]{Illustration of the underestimation of the errors. The contours of constant $\chi^2$ near the minimum of an $R^{1/4}$ fit to a $h/R_{eB}=0.5$, $D/B=0.1$ disk plus bulge model. The cross shows the best-fit solution, the circle near the upper left corner gives the real parameters of the model. The errors estimated at the $5\sigma$ contour underestimate the differences between the model and the fit by a factor 2 (see \S \ref{parameter}).} \label{figcontour} \end{figure} \subsection{Sky subtraction errors} \label{sky} Sky subtraction errors can induce severe systematic errors on the derived photometric parameters of galaxies. Figure \ref{figskyerror} shows the parameters derived from the $R^{1/4}$ plus exponential models examined in the previous section, where now the sky has been overestimated or underestimated by $\pm 1$\%. The sky correction algorithm is not activated. The biases become increasingly large as the sky brightness approaches the effective surface brightness of the models. As expected, underestimating the sky (a negative sky error) produces total magnitudes that are too bright and half-luminosity radii that are too large relative to the true ones. The size of the bias correlates with the extrapolation needed to derive $M_{TOT}^f$. The opposite happens when the sky is overestimated, but the amplitude of the bias is smaller, because there is no extrapolation. The correlated error $\Delta FP$ remains small ($\approx 0.05$), except for the cases where large extrapolations are involved. The D/B ratio is ill determined, with better precision for models with extended disks ($h/R_{eB}>2.5$). The scale length of the bulge is better determined for low values of $D/B$ (dominant bulge), the scale length of the disk component is better determined for large values of $D/B$ (dominant disk). The parameter least affected is the value $\Gamma$ of the seeing, which is determined in the inner, bright parts of the models, where sky subtraction errors are unimportant. Bulge-only or disk-only models appear to be fit best by two-component models (crosses and triangular crosses in Figure \ref{figskyerror}). Finally, note that reasonably good fits ($\chi^2<10$) to the surface brightness profiles are always obtained, in spite of the large errors on the reconstructed parameters. The biases discussed above can be fully corrected when the sky-fitting algorithm of Eq. \ref{chidmax} is applied. Figure \ref{figskyfix} shows the reconstructed parameters of the models considered in \S \ref{parameter}, where sky subtraction errors of 0, $\pm1$\%, $\pm3$\% have been introduced. For most of the models examined, the errors on the derived quantities are no more than a factor 2 larger than those shown in Fig. \ref{figresdb}. The sky corrections are computed to better than 0.5\% precision. Larger errors $\Delta M_{TOT}$ and $\Delta R_e$ are obtained for models with relatively weak ($D/B<0.3$) and extended disks ($h/R_{eB}>2.5$), where the degeneracy discussed in \S \ref{parameter} is complicated by the sky subtraction correction. These cases give reasonably good fits ($\chi^2<10$), but are identified by the large extrapolation ($>0.3$) involved. Models with concentrated disks ($h/R_{eB}<0.2$) can also be difficult to reconstruct, when $h/\Gamma\approx 1$. For some of these problematic fits, one-component solutions are preferred by Eqs. \ref{chibo}-\ref{chido} (circles in Eq. \ref{figskyfix}). \begin{figure} \plotone{f6.eps} \caption[f6.eps]{The biases introduced by a $\pm 1$\% sky subtraction error. Quantities plotted as in Fig. \ref{figresdb}. Models with $D/B=0$ which have been fitted with two components are shown as crosses. Models with $B/D=0$ which have been fitted with two components are shown as triangular crosses. Note the change of scale on the ordinate axis with respect to Figure \ref{figresdb}. See discussion in \S \ref{sky}.} \label{figskyerror} \end{figure} \begin{figure} \plotone{f7.eps} \caption[f7.eps]{The effects of the sky fitting algorithm. The parameters of the models of Fig. \ref{figskyerror} with the sky subtraction errors of 0, $\pm1$\%, $\pm3$\%, are reconstructed using the sky fitting algorithm. Quantities and symbols plotted as in Figures \ref{figresdb} and \ref{figskyerror}. In addition, the difference $\Delta$dSky=dSky/Sky-dSky$^f/$Sky on the sky correction is plotted. Note the change of scale on the ordinate axis with respect to Figure \ref{figresdb}. See discussion in \S \ref{sky}.} \label{figskyfix} \end{figure} A common problem of CCD galaxy photometry is the relatively small field of view, particularly with the older smaller CCDs. If the size (projected on the sky) of the CCD is not large enough compared to the half-luminosity radius of the imaged galaxy, then the sky as determined on the same frame will be contaminated by galaxy light and biased to values larger than the true one. Total magnitudes and half-luminosity radii can therefore be biased to smaller values, the effect being more important for intrinsically large galaxies, which tend to have low effective surface brightnesses. The mean surface brightness in the annulus with radii $R_i^{max}$ and $2R_i^{max}$ (see \S \ref{diskbulge}) predicted by the fit allows us to estimate the size of the contamination. \subsection{Radial extent} \label{extension} Photoelectric photometry of large, nearby galaxies rarely goes beyond 1 or 2 $R_e$ (Burstein et al. 1987) and the same applies for the surface photometry obtained with smallish CCDs. The typical profiles obtained in Paper III extend to a least 4 $R_e$, but a small fraction of them are less deep, reaching 1 or 2 $R_e$ only. Here we investigate the effect of the radial extent of the profiles, keeping the normalization of the profiles fixed ($10^7$ counts, $S/N\approx 10^3$). Sky subtraction errors of $0$, $\pm3$\% are introduced and the sky fitting is activated. Figure \ref{figextension} shows the cumulative distributions of the errors on the derived photometric parameters as derived from the simulations, for a range of $R_{max}$ values. When $R_{max}=R_e$, rather large errors are possible (0.3 mag in the total magnitude, $>30$\% in $R_e^f$). The main source of error is again the large extrapolation involved when $R_{max}\approx R_e$, coupled with the sky correction which becomes unreliable for these short radial extents. As soon as $R_{max}\ge3R_e$ the errors reduce to the ones discussed in \S \ref{parameter}. The same kind of trend is observed for the parameters of the two component ($\Delta (D/B)$, $\Delta R_{eB}$, $\Delta h$). The seeing values are less affected, as they are sensitive to the central parts of the profiles only. Finally, note that in all cases very good fits are obtained ($\chi^2\approx 1$). \begin{figure} \plotone{f8.eps} \caption[f8.eps]{The effect of the radial extent of the profiles on the precision of the derived parameters. The cumulative distributions of the errors on the derived photometric parameters as derived from the simulations are shown for a range of $R_{max}$ values (full lines: $R_{max}=R_e$, dotted lines: $R_{max}=2R_e$, dashed lines $R_{max}=3R_e$, long-dashed lines: $R_{max}=4R_e$). Good reconstructions are obtained when $R_{max}/R_e>2$ (see \S \ref{extension}).} \label{figextension} \end{figure} \subsection{Signal-to-noise ratio} \label{snratio} For most of the galaxies discussed in Paper III, multiple profiles are available with integrated signal-to-noise ratios $S/N> 300$, the normalization used in the previous sections. But for some of the luminosity profiles a smaller number of total counts has been collected (see Figure \ref{figparspace}). Here we investigate how the signal-to-noise ratio of the profiles affects the outcome of the fits. As before, the subset of models of \S \ref{sky} is used with $R_{max}\le4 R_e$ (see comment at the beginning of \S \ref{montecarlo}). Sky subtraction errors of $0$, $\pm3$\% are introduced and the sky fitting is activated. Figure \ref{figflux} shows how the errors on the derived parameters increase when the signal-to-noise ratio is reduced. For fluxes as low as about $10^5$ ($S/N\approx 10^2$) all of the derived photometric parameters become uncertain (0.2 mag in the total magnitudes, 20\% variations in the derived $R_e$, large spread $\Delta (D/B)$, $\Delta R_{eB}$, $\Delta h$, $\Delta \Gamma$), as large extrapolations and uncertain sky corrections are applied. In all cases very good fits are obtained ($\chi^2\approx 1$). \begin{figure} \plotone{f9.eps} \caption[f9.eps]{The effect of the signal-to-noise ratio of the profiles on the precision of the derived parameters. Good reconstructions are obtained when $S/N>300$ (see \S \ref{snratio}). Note the change of scale on the ordinate axis with respect to Figure \ref{figresdb}.} \label{figflux} \end{figure} \subsection{Seeing and sampling effects} \label{seeing} Some of the galaxies considered in Paper III are rather small, with $R_e<4''$. Here we investigate the effects of seeing and pixel sampling, when $R_e\approx\Gamma\approx$pixel size. Figure \ref{figseeing} shows that reliable parameters can be derived down to $R_e\approx\Gamma$, with pixel sizes 0.4-0.8 arcsec, with only a small increase of the scatter for $R_e<2\Gamma$. A small systematic effect is caused by the choice of the psf. Saglia et al. (1993) demonstrate that a good approximation of the psfs observed during the runs described in Paper III is given by the $\gamma$ psf with $\gamma=1.5-1.7$. We adopt $\gamma=1.6$ for the fits. Here we test the effect of having $\gamma=1.5$ or 1.7 with a pixel size of 0.8 arcsec. Figure \ref{figgamma} shows that if $\gamma=1.5$ is the true psf of the observations, then the half-luminosity radius, the total luminosity, the scale length of the bulge will be slightly overestimated, and the disk to bulge ratio slightly underestimated. A small systematic trend is observed in the correlated errors $\Delta FP$. The scale length of the disk component is less affected. The sky corrections are also biased, but do not strongly affect the photometric parameters, because of the high average surface brightness of the small $R_e$ models. Seeing values suffer a very small, but systematic effect. The opposite trends are observed if the true $\gamma$ is 1.7. In all cases very good fits are obtained ($\chi^2\approx 1$). The systematic differences become unimportant for $R_e>2\Gamma$. Finally, the seeing values derived can be systematically biased, if the central concentration of the fitted galaxies does not match the one of the $R^{1/4}$ plus exponential models. We investigate this effect by fitting the $\Psi=12$ plus exponential or the smoothed $R^{1/4}$ plus exponential models discussed in \S \ref{diskbulge}. We find that in the first case the seeing value is underestimated which compensates for the higher concentration of the $\Psi=12$ component. The shallow radial decline of the luminosity profile in the outer parts introduces systematic biases in the reconstructed parameters, similar to those discussed for the $R^{1/n}$ profiles, for large values of $n$ (see \S \ref{profiles}). The half-luminosity radii and total magnitudes derived are underestimated by 20\% and 0.2 mag respectively, when a $\Psi=12$ model with no exponential component is fitted. The biases are reduced when models with an exponential component are constructed. In the case of the smoothed $R^{1/4}$ law, the seeing value is overestimated to fit the lower concentration of the smoothed $R^{1/4}$ component. No biases are introduced on the other reconstructed parameters. \begin{figure} \plotone{f10.eps} \caption[f10.eps]{The effect of seeing and pixel sampling of the profiles on the precision of the derived parameters. Different symbols indicate different pixel sizes (small dot 0.4 arcsec, triangles 0.6 arcsec, squares 0.8 arcsec). Note the expanded ordinate scale with respect to Figures \ref{figquatexp}-\ref{figflux}. See discussion in \S \ref{seeing}.} \label{figseeing} \end{figure} \begin{figure} \plotone{f11.eps} \caption[f11.eps]{The effect of the choice of the psf on the precision of the derived parameters. Open triangles for $\gamma=1.5$, dots for $\gamma=1.6$ and open squares for $\gamma=1.7$. Fits performed with the $\gamma=1.6$ psf overestimate (underestimate) magnitudes and half-luminosity radii of models constructed with $\gamma=1.5$ ($\gamma=1.7$; see \S \ref{seeing}). Note the expanded ordinate scale with respect to Figures \ref{figquatexp}-\ref{figflux}.} \label{figgamma} \end{figure} \subsection{Tests of profile combination} \label{testcombination} In order to test the combination algorithm described in \ref{combination}, four profiles with different $\Gamma$, pixel sizes, normalizations, gain, readout noise, and sky subtraction errors (see Table \ref{tabcombination}; these parameters match the typical values of the profiles of Paper III) are generated for the set of models identified by the open pentagons of Figure \ref{figparspace}. Figure \ref{figcombination} shows the result of the test. The abscissa plots the residuals $\Delta$ of the parameters derived using the fitting procedure with profile combination. $\Delta$ dSky and $\Delta \Gamma$ are averaged over the four obtained values. The ordinate plots the {\it mean} of the residuals of the parameters derived by fitting each single independently as crosses, and the residuals of each fit as dots. The profile combination algorithm obtains better precision on all of the parameters with the exception of $\Gamma$, where the maximum deviation is in any case smaller than 8\%. \begin{figure} \plotone{f12.eps} \caption[f12.eps]{The profile combination algorithm and the precision of the derived parameters. The x-axis plots the residuals $\Delta$ of the parameters derived using the fitting procedure with profile combination. $\Delta$ dSky and $\Delta \Gamma$ are averaged over the four obtained values. The y-axis plots the {\it mean} of the residuals of the parameters derived by fitting each single independently as crosses, and the residuals of each fit as dots (see discussion in \S \ref{testcombination}).} \label{figcombination} \end{figure} \clearpage \begin{table}[ph] \centering \caption[*]{The parameters of the multiple profiles test (see \S \ref{testcombination}).} \label{tabcombination} \begin{tabular}{lrrrrr} &&&&\\ \tableline \tableline Profile & 1 & 2 & 3 & 4 \\ \tableline &&&&\\ Pixel size $('')$ & 0.4 & 0.606 & 0.862 & 0.792 \\ Sky per pixel & 300 & 350 & 250 & 1500 \\ $\delta$Sky/Sky & $+1$\% & $-0.5$\% & $+1.5$\% & $+0.5$\% \\ $R_{max}/R_e$ & 4 & 3 & 4.5 & 2.5 \\ Normalization & $10^7$ & $5\times 10^6$ & $10^7$ & $5\times 10^6$ \\ Gain & 1 & 3 & 1 & 2 \\ Ron & 1 & 4 & 1 & 5 \\ $\Gamma ('')$ & 2 & 2.1 & 1.5 & 2.4 \\ \tableline \end{tabular} \end{table} \subsection{``Bulge'' and ``Disk'' components} \label{decomposition} The discussion of the previous sections shows that for a large fraction of the parameter space, i.e. when deep enough profiles are available, with large enough objects, not only can the global photometric parameters $R_e$ and $M_{TOT}$ be reconstructed with high accuracy, but also the parameters of the $R^{1/4}$ and the exponential components. Here we investigate further if reliable ``bulge'' and ``disk'' parameters can be derived, when the profiles analysed are constructed from the superposition of these two components. With this purpose, we constructed a number of two-dimensional frames (filled triangles in Figure \ref{figparspace}) as the sum of a flattened $R^{1/4}$ bulge and an exponential disk of given inclination. The bulge (disk) frames follow an exact $R^{1/4}$ (exponential) law with $R_{eB}=12\sqrt{b/a}$ arcsec ($h=10\sqrt{\cos(i)}$ arcsec) along the minor axis. Three flattenings of the bulge ($b/a=1,0.7,0.4$), four inclinations for the disk ($i=0^\circ,30^\circ,60^\circ,80^\circ$, where $i=0^\circ$ is face-on and $i=90^\circ$ edge-on) and five values of the disk to bulge ratio ($D/B=0,0.5,1,2,\infty$) are considered. The resulting models are normalized to $10^7$ counts. The pixel size is 0.6 arcsec. The circularly averaged luminosity profiles are derived following the same procedure adopted for the observed galaxies (see Paper III) and extend out to $\approx 4-6R_e$. A 1\% sky error is introduced and the sky fitting procedure is activated. Note that the maximum flattening of the EFAR galaxies is $b/a=0.5$, with 96\% of the galaxies having $b/a>0.6$ (see Paper III). This corresponds to (pure) disk inclinations $i\le 60^\circ$. Figure \ref{figbulgedisk} shows the reconstructed parameters as a function of the inclination angle of the disk, for the different flattenings of the bulge, using the sky fitting procedure. The horizontal bars show models with $D/B=0.5$. The plot at the bottom right shows the scale lengths of the flattened bulge (filled symbols) or of the inclined disk as a function of the flattening angle (open symbols, $i=arccos(b/a)$) or of the inclination angle, normalized to the $b/a=1$ or $i=0^\circ$ values. When $D/B$ is low ($\le 0.5$) the errors are very small for {\it every} inclination angle. For larger values of $D/B$, reliable photometric parameters are obtained for $i<60^\circ$, but as soon as the disk is nearly edge-on, total magnitudes and half-luminosity radii are overestimated (by 0.1 mag and 20\% respectively). The integrated circularized profiles, in fact, converge more slowly than the ones following the isophotes. The correlated errors $\Delta FP$ always remain very small. Similarly, the parameters of the two components are reconstructed well for $i<60^\circ$, but badly underestimate the disk when it is nearly edge-on. However, a decent fit is obtained, by increasing the half-luminosity radius of the bulge component (see Fig. \ref{figbadfit} (e) and (f)). The sky correction is returned to better than 0.5\% for $i<80^\circ$. The systematic effects connected to the flattening of the bulge are small for the range of ellipticities considered here ($b/a\ge0.6$). These results indicate two potential problems, (i) galaxies may be misclassified due to the presence of an edge-on disk component not being recognized, or (ii) the photometric parameters may be systematically overestimated. However, these problems do not apply to the EFAR sample, where $b/a>0.5$ always and $b/a\ge 0.6$ for 96\% of the galaxies. Therefore, galaxies with bright edge-on disks are only a very small fraction. Galaxies with faint edge-on disks, which may not show large averaged flattenings, have low $D/B$ ratios and therefore are not affected by problem (ii). In a future paper we will address the question whether in these cases the isophote shape analysis might detect these faints disks and improve on point (i). Finally, the two-dimensional frames described here have been used to calibrate the estimator of the galaxy light contamination described in \S \ref{diskbulge}. We measured the sky in the same way as for the real frames of Paper III, by considering some small areas around the simulated galaxies. We find that the predicted galaxy light contamination overestimates the measured sky excess by at least a factor two, and therefore can be used as a rather robust upper limit to the galaxy light contamination. \begin{figure} \plotone{f13.eps} \caption[f13.eps]{The reconstructed parameters of the bulge plus disk models as a function of the inclination $i$ of the disk. Different symbols indicate different flattenings of the bulge. The horizontal bars show models with $D/B=0.5$. The plot at the bottom right shows the scale lengths of the flattened bulge (open symbols) or of the inclined disk (filled symbols) as a function of the flattening angle ($i=arccos(b/a)$) or of the inclination angle, normalized to the $b/a=1$ or $i=0^\circ$ values. Good reconstructions of the parameters are obtained when the inclination is less than $60^\circ$ (see \S \ref{decomposition}).} \label{figbulgedisk} \end{figure} \subsection{$R^{1/n}$ luminosity profiles} \label{profiles} The tests described above show that our fitting algorithm is able to reconstruct the parameters of a sum of an $R^{1/4}$ plus an exponential law accurately. In these cases sky subtraction errors can also be corrected efficiently. Even so, we do find in Paper III that luminosity profiles of real early-type galaxies show systematic differences from $R^{1/4}$ plus exponential profiles, yielding to a median reduced $\chi^2$ of 6. Here we quantify the systematic effects that would be produced in this case, by studying the case of the $R^{1/n}$ profiles. CCO fitted the luminosity profiles of 52 early-type galaxies using the $R^{1/n}$ law introduced by Sersic (1968): \begin{equation} \label{r1n} I(R)=I_e^n 10^{-b_n\left[\left(\frac{R}{R_e^n}\right )^{1/n}-1\right]}, \end{equation} where $b_n\approx 0.868n-0.142$, $R_e^n$ is the half-luminosity radius, and $I_e^n$ the surface brightness at $R_e^n$. The total luminosity is $L_T=K_nI_e^n{R_e^n}^2$, where $\log K_n\approx 0.03[\log(n)]^2+0.441\log(n) +1.079$. Eq. \ref{r1n} reduces to Eq. \ref{bulge} for $n=4$ and to Eq. \ref{disk} for $n=1$. For large values of $n$, Eq. \ref{r1n} describes a luminosity profile which is very peaked near the center and has a very shallow decline in the outer parts. Ciotti (1991) computes the curve of growth related to Eq. \ref{r1n} analytically for integer values of $n$ and finds that while already $\approx 13$\% of the total light is included inside $R<0.05 R_e^n$, only 80\% of the total light is included inside $6R_e^n$ for $n=10$. We fitted Eq. \ref{r1n}, modified to have a core at $R<0.05 R_e^n$, to an $R^{1/4}$ plus exponential model for $n=0.5$ to $n=15$ out to $6R_e^n$. Fig. \ref{figr1nfit} shows the results of the fit for a selection of models. With the exception of the $n=0.5$ model, all of the $R^{1/n}$ profiles can be described by a combination of an $R^{1/4}$ and an exponential component, with residuals less than 0.2 mag arcsec$^{-2}$ for $R\le 4R_e$. For $n<4$ the residuals increase to 0.4 mag arcsec$^{-2}$ at $R>5R_e^n$, where the fits are increasingly brighter than the $R^{1/n}$ profiles. For large values of $n$ the residuals reach -0.4 mag arcsec$^{-2}$ at $R>5 R_e^n$, where the fits are increasingly fainter than the $R^{1/n}$ profiles. The relation between $n$ and the parameters of the decomposition is shown in Fig. \ref{figr1npar}. Models with $1<n<4$ are fitted using a decreasing amount of the exponential component, with a scale length comparable to the one of the $R^{1/4}$ component. Models with $n>4$ are fitted with an increasing amount of the exponential component, with increasingly large scale length. Half-luminosity radii are progressively underestimated, being $\approx 60$\% of the true values at $n=15$. Correspondingly, total magnitudes are also underestimated, by 0.25 magnitudes at $n=15$. A possible problem can emerge for large values of $n$, if the sky fitting algorithm is activated. The dotted curves in Fig. \ref{figr1npar} show that if the sky subtraction algorithm is activated (Eq. \ref{chidmax}), then larger systematic effects are produced. Note that the computed sky correction (dotted curve of Fig. \ref{figr1npar}) is $\approx 0$\ for $n \approx 1$ or $n\approx 4$ only. For $n>4$ the correction is used to reduce the systematic negative differences in the outer parts of the profiles. A comparison between the fitted sky corrections and the upper limits on the possible galaxy light contamination (see \S \ref{diskbulge} and \ref{decomposition}) gives an important consistency check. In the case shown in Fig. \ref{figr1npar} the fitted sky corrections are twice as large as the upper limits on the galaxy light contamination. In a real case this, together with the rather large values of $\chi^2$, would hint at an uncertain fitted sky correction. The fact that the $R^{1/n}$ sequence can be approximated by a subsample of $R^{1/4}$ plus exponential models suggests a possible reinterpretation of CCO's results: the variety of profile shapes of early-type galaxies is caused by the presence of a disk component. Moreover, the use of the $R^{1/n}$ profiles to determine the photometric parameters of galaxies of large $n$ is dangerous, since the extrapolation involved is large and the fitted profiles barely reach 2 or 3$R_e^n$, as derived from the fit. This problem is much smaller using the $R^{1/4}$ plus exponential approach. \begin{figure} \plotone{f14.eps} \caption[f14.eps]{The fits to the $R^{1/n}$ law. Two plots are drawn for each value of $n$ (given in the top right corner). In the top plot the crosses (one point in every seven) show the luminosity profiles $\mu(R)=-2.5 \log I(R)$ of the $R^{1/n}$ law as a function of $R/R_e$. The dotted and dashed curves show the best-fitting $R^{1/4}$ and exponential laws respectively. In the bottom plot the residuals (full curves) in mag arcsec$^{-2}$ from the fits to the $R^{1/n}$ law are shown. The dashed curve shows the residuals (in mag) from the curves of growth (see discussion in \S \ref{profiles}).} \label{figr1nfit} \end{figure} \begin{figure} \plotone{f15.eps} \caption[f15.eps]{The relation between $n$ and the parameters of the decomposition (see \S \ref{profiles}). The full curves refer to the results obtained with no sky subtraction errors. The dotted curves show the results obtained when the sky fitting algorithm is activated.} \label{figr1npar} \end{figure} \subsection{Discussion} \label{discussion} We conclude our tests by discussing the quality parameters defined in Table \ref{tabquality} and their use to estimate the size of the systematic errors present. The definitions given in Table \ref{tabquality} have been derived after inspection of Figures \ref{figresdb} to \ref{figr1npar}, with the desired goal of identifying three classes of precision, $\Delta M_{TOT}\le 0.05$, $\Delta M_{TOT}\le 0.15$, $\Delta M_{TOT}>0.15$. The parameters $Q_E$, $Q_{max}$, $Q_{\chi^2}$, $Q_{S/N}$, and $Q_\Gamma$ are directly related to the simulations. Their low values imply that the fits involve a small extrapolation, extend to large enough radii, give low surface brightness residuals with a large enough signal-to-noise ratio and good spatial resolution. The definitions of $Q_{\hbox{Sky}}$ and $Q_{\delta \hbox{Sky}}$ deal with the accuracy of the sky subtraction, taking into account that high surface brightness galaxies suffer less from this problem, and that large sky corrections indicate a lower quality of the data. Low values of $Q$ (see Eq. \ref{qtot}) imply low values of all quality parameters. Figure \ref{figqualdis} shows the cumulative distributions of the errors $\Delta M_{TOT}$, $\Delta R_e$ and $\Delta FP$ derived from all the performed disk plus bulge fits with sky correction algorithm activated, as a function of the different quality parameters. The two most important parameters regulating the precision of the photometric parameters are the level of extrapolation and the goodness of the fit, followed by the sky subtraction errors. A low $Q_E$ fixes the maximum possible overestimate of the parameters. A low $Q_{\chi^2}$ with a low $Q_E$ constrains the underestimate and the reliability of the sky correction. The ranges of the errors match the desired goal of identify three classes of precisions. Finally, it is sobering to note that the constraints needed to achieve $Q=1$, high precision total magnitudes and $R_e$ are rather stringent. Only 16\% of EFAR galaxies have $Q=1$. Most of the existing {\it published} values of $M_{TOT}$ and $R_e$ of galaxies are far below this precision, because of the restricted radial range probed by photoelectric measurements or small CCD chips, because of sky subtraction errors, and also by the use of the pure $R^{1/4}$ curve of growth fitting (see Figure \ref{figquatexp}). The related observational problems can be somewhat reduced with the use of large CCDs (see Introduction), but the {\it a priori} limiting factor of galaxy photometry, the extrapolation, will always remain with us at a certain level. On the other hand, the errors on $M_{TOT}$ and $R_e$ are strongly correlated, so that the quantity $\log R_e-0.3\langle SB_e \rangle$ is always well determined. This fact allows the accurate distance determinations achieved using the Fundamental Plane correlations despite the systematic errors in the photometric quantities. \begin{figure} \plotone{f16.eps} \caption[f16.eps]{The precision of the reconstructed total magnitudes $M_{TOT}$, the half-luminosity radii $R_e$ and the combined quantity $FP=\log R_e - 0.3 \langle SB_e\rangle $. The cumulative distributions of the errors $\Delta M_{TOT}$, $\Delta R_e$ and $\Delta FP$ derived from all the performed disk plus bulge fits with sky correction algorithm activated are shown as a function of the different quality parameters defined in \S \ref{quality}. The full lines plot the distributions when the parameters have value of 1, the dotted ones when the value is 2, the dashed ones when the value is 3. The distributions derived by selecting on the global quality parameter $Q$ match the precision ranges identified in \S \ref{discussion}.} \label{figqualdis} \end{figure} \section{Conclusions} \label{conclusions} We constructed an algorithm to fit the circularized profiles of the (early-type) galaxies of the EFAR project, using a sum of a seeing-convolved $R^{1/4}$ and an exponential law. This choice allows us to fit the large variety of profiles exhibited by the EFAR galaxies homogeneously. The procedure provides for an optimal combination of multiple profiles. A sky fitting option has been developed. A conservative upper limit to the sky contamination due to the light of the outer parts of the galaxies is estimated. From the tests described in previous sections we draw the following conclusions: 1) The reconstruction algorithm applied to simulated $R^{1/4}$ plus exponential profiles shows that random errors are negligible if the total signal-to-noise ratio of the profiles exceeds $300$. Systematic errors due to the radial extent of the profiles are minimal if $R_{max}/R_e> 2$. Systematic errors due to sky subtraction are significant (easily larger than 0.2 mag in the total magnitude) when the sky surface brightness is of the order of the average effective surface brightness of the galaxy. They can be reliably corrected for as long as the fitted profiles show small systematic deviations ($\chi^2<12.5$). 2) Strong systematic biases (errors larger than 0.2 mag in the total magnitudes) are present when a simple $R^{1/4}$ or exponential model is used to fit test profiles with disk to bulge ratios as low as 0.2. 3) The use of the shape of the (normalized) $\chi^2$ function badly underestimates the (systematic) errors on the photometric parameters. 4) Systematic biases emerge when test profiles are derived for systems with significant disk components seen nearly edge-on, or when the fitted luminosity profile declines more slowly than an $R^{1/4}$ law. The parameters of bulge plus disk systems can be determined to better than $\approx 20$\% if the disk is not very inclined ($i<60^\circ$). 5) The sequence of $R^{1/n}$ profiles, recently used to fit the profiles of elliptical galaxies by Caon et al. (1993), is equivalent to a subset of $R^{1/4}$ and exponential profiles, with appropriate scale lengths and disk-to-bulge ratios, with moderate systematic biases for $n\le 8$ and residuals less than 0.2 mag arcsec$^{-2}$ for $R\le 4R_e$. This suggests that the variety of luminosity profiles shown by early-type galaxies is due to the frequent presence of a weak disk component. 6) A set of quality parameters has been defined to control the precision of the estimated photometric parameters. They take into account the amount of extrapolation involved to derive the total magnitudes, the size of the sky correction, the average surface brightness of the galaxy relative to the sky, the radial extent of the profile, its signal-to-noise ratio, the seeing value and the reduced $\chi^2$ of the fit. These are combined into a single quality parameter $Q$ which correlates with the expected precision of the fits. Errors in total magnitudes $M_{TOT}$ less than 0.05 mag and in half-luminosity radii $R_e$ less than 10\% are expected if $Q=1$, and less than 0.15 mag and 25\% if $Q=2$. 89\% of the EFAR galaxies have fits with $Q=1$ or $Q=2$. The errors on the combined Fundamental Plane quantity $FP=\log R_e -0.3\langle SB_e\rangle$, where $\langle SB_e \rangle$ is the average effective surface brightness, are smaller than 0.03 even if $Q=3$. Thus systematic errors on $M_{TOT}$ and $R_e$ only marginally affect the distance estimates which involve $FP$. \acknowledgments {RPS acknowledges the support by DFG grants SFB 318 and SFB 375. GW is grateful to the SERC and Wadham College for a year's stay in Oxford, and to the Alexander von Humboldt-Stiftung for making possible a visit to the Ruhr-Universit\"at in Bochum. MMC acknowledges the support of a Lindemann Fellowship, a DIST Collaborative Research Grant and an Australian Academy of Science/Royal Society Exchange Program Fellowship. This work was partially supported by NSF Grant AST90-16930 to DB, AST90-17048 and AST93-47714 to GW, AST90-20864 to RKM, and NASA grant NAG5-2816 to EB. The entire collaboration benefitted from NATO Collaborative Research Grant 900159 and from the hospitality and monetary support of Dartmouth College, Oxford University, the University of Durham and Arizona State University. Support was also received from PPARC visitors grants to Oxford and Durham Universities and a PPARC rolling grant: ``Extragalactic Astronomy and Cosmology in Durham 1994-98''.}
proofpile-arXiv_065-487
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{ Introduction} The well-studied quasicrystals fall essentially into two families, the Al-transition metal family exemplified by i(AlPdMn) and the Frank-Kasper family \cite{FK1,FK2,shoe57,shoe87} exemplified by i(AlCuLi). Decagonal quasicrystals are known experimentally only in the first family, except for one claim \cite{chinese}. Recently, three dimensional model systems have been found which, in simulations, freeze into a quasicrystal phase with Frank-Kasper local structure: a monatomic system with a special potential freezes into a dodecagonal quasicrystal \cite{dzug}, and a system of large (L) and small (S) atoms with Lennard-Jones interactions freezes into an icosahedral phase \cite{rotcool}, in fact a simplified version of the Henley-Elser structure model for i(AlCuLi) and i(AlZnMg) \cite{henels}. Since Frank-Kasper systems are dominated by $sp$ conduction electrons, it is believed that simple pair potentials (as used in these simulations) can offer a good approximation to the structure energy \cite{haf}. The L and S atoms used by Roth et al.\ \citeyear{rotcool} had single-well potentials where the ideal distances are non-additive $(r_{LL} \approx r_{LS} \approx 1.15 r_{SS})$. This paper reports on a new {\it decagonal} Frank-Kasper phase which froze from the liquid in molecular dynamics simulations of a model system using practically the same potentials as Roth et al.\ \citeyear{rotcool} (see Sec.~\ref{simulations}). No realization of this structure in nature has been established (except conceivably in d(FeNb), as discussed in Sec.~\ref{comparison}). However, in the absence of a {\it realistic} model system, a toy model system with an equilibrium quasicrystal phase can be quite useful. (To date, microscopically derived potentials for particular alloy systems have not been used in simulations of freezing from the melt \cite{mihII}, but only for comparisons of trial structures at zero temperature, in the Al-transition metal case (Phillips, Deng, Carlsson and Murray, 1991; Zou and Carlsson, 1994; Widom, Phillips, Zou and Carlsson, 1995; Widom and Phillips, 1993; Phillips and Widom, 1994; Mihalcovi{\v c}, Zhu, Henley and Oxborrow, 1996; Mihalcovi{\v c} et al.\ 1996b) \nocite {carlsson1,carlsson2,carlsson3,wid92,phil,mihI,mihII}). Of course, there is some fundamental interest in the existence of any microscopic model of interacting atoms which has a quasicrystal equilibrium state. Furthermore, within such a model one can study the phonon/phason coupling, the energy changes and barriers corresponding to tile flips, the phason elastic constants with their temperature dependence, and can derive from microscopics an effective Hamiltonian in terms of tile-tile interactions (Mihalcovi{\v c} et al.\ 1996a). In two dimensions, the ``binary tiling'' case \cite{lan86,wid87} has served this purpose, but it has quite unrealistic bond radii, in contrast to our model potentials (which were in fact tailored to favor the i(AlCuLi) structure \cite{rotstab}). The bulk of our paper presents an idealized decoration model which was constructed by abstracting the patterns observed in our simulations. The model has several variants, corresponding to closely related atomic structures, yet described by different tilings (See Sec.~\ref{dsdm}). Thus it exemplifies a system where an effective tile-tile Hamiltonian may select a less disordered subensemble of a random tiling ensemble \cite{jeong}. We have constructed the acceptance domains of a quasiperiodic version of our model (Sec.~\ref{environments}). Finally, by applying information from the ideal model, we adjusted the potentials so as to obtain a better-ordered structure after quenching (Sec.~\ref{XXX}). In the concluding section, we have discussed how our model is related to others, and how realistically it extends beyond our small system size. \section { Initial Simulations} \label{simulations} Our decagonal phase was first seen in a slow-cooling simulation of a binary liquid. (Further details about this simulation and other structures observed, as well as the motivation for our initial choice of parameters for the potential, can be found (Roth et al.\ 1995). \nocite{rotcool} The interaction is described by Lennard-Jones potentials \begin{equation} v_{\alpha \beta}(r) = 4\epsilon_{\alpha \beta} \left(\left(\frac{\sigma_{\alpha \beta}}{r}\right)^{12} - \left(\frac{\sigma_{\alpha \beta}}{r}\right)^{6}\right) \label{eq-LJ} \end{equation} The bond parameters $\sigma_{\alpha \beta}$ are $\sigma_{SS} = 1.05 $, $\sigma_{SL} = 1.23 $, and $\sigma_{LL} = 1.21 $. These and all other simulation parameters are given in reduced or Lennard-Jones units, and are indicated by a star. The corresponding bond lengths are $2^{1/6} \sigma_{\alpha\beta}$. In overall outline, these distances are appropriate for a typical tetrahedrally-close-packed structure, in which the small atoms have icosahedral coordination shells and the large atoms have Frank-Kasper coordination shells with larger coordination number. (Frank-Kasper structures are by definition ``tetrahedrally close-packed (tcp)'', meaning that neighboring atoms always form tetrahedra (Samson, 1968; Samson, 1969; Shoemaker et al.\ 1957; Frank and Kasper, 1958; Frank and Kasper, 1959)\nocite {sam1,sam2,shoe57,FK1,FK2}. This is not strictly true in our model -- or that of Henley and Elser (1986) -- since there is a small number of non-tcp environments.) The coupling constants were set to $\epsilon_{SS} = \epsilon_{LL} = 0.656$ for same species and to $\epsilon_{SL} = 1.312$ for different species, in order to prevent phase separations into monatomic domains. To save computation time we cut off the potential smoothly at $r_{c} = 2.5 \sigma_{SL}$. For the simulation we have modified the standard molecular dynamics method to allow us to control pressure and temperature as described by Evans \& Morriss (1983). \nocite{eva} A constant cooling rate may be introduced by the method described by Lan\c{c}on \& Billard (1990). \nocite{lanbil} The equations of motion are integrated in a fourth-order Gear-predictor-corrector-algorithm (see for example Allen \& Tildesley (1987)). The time increment $\delta t^{*}$ is adjusted after testing for numerical stability. We find that $\delta t^{*} = 0.00462$ is an appropriate value. For simplicity the masses of small and large atoms are set to unity. Our simulation box is a cube containing 128 small (S) and 40 large (L) atoms. The initial positions are set with a random number generator, and then relaxed. The generated liquid is equilibrated at an initial temperature of about $T^{*} = 1.0$. At high cooling rates we observe a transition form a supercooled liquid to a glass at about $T^* = 0.5$. If the cooling rate is lowered to $T^{*}/t^{*} = 1.74\times 10^{-5}$ the supercooled liquid transforms sharply to an ordered solid at about $T^* = 0.6$, which analysis reveals is a quasicrystal (sometimes icosahedral and sometimes decagonal). The solid may still contain defects, but the order can be improved by annealing. The annealed structures are quenched down to $T^* = 0$ for structure analysis. Subsequent annealing runs with $5 \times 10^{5}$ steps at constant temperatures of $T^* = 0.5 $ and 0.4 show that the nucleated structures are stable in the sense that they do not change except for defect annealing. More details about the cooling simulations can be found by Roth et al.\ \citeyear{rotcool}. The new decagonal structure was observed at cooling rates of $T^{*}/t^{*} = 1.74\times 10^{-6}$ and $1.74\times 10^{-5}$. The tenfold axis is aligned with the $\langle 1 1 0 \rangle$ diagonal of the cubic simulation cell; the structure is stacked periodically, with a stacking vector parallel to this direction. Fig.~\ref{fig-proj} shows a projection of the simulation box onto a plane normal to this axis; the atom positions are highly ordered and the approximant tenfold symmetry within this plane is obvious. Because of the constraint of fitting into the fixed, periodic simulation cell, the structure is distorted somewhat (in that the $x/y$ ratio within the layers is squeezed relative to decagonal symmetry). By examining slices transverse to the decagonal axis (not shown), we could distinguish three different types of layers transverse to the tenfold axis, which we label ``A'', ``B'', and ``X''. Let $c$ be the lattice constant in the stacking direction; the simulation box extends over two periods ($2c$) before encountering the same atomic layer. Most of the atoms in our structure are in A and B layers, which are separated by $\sim 0.8$ units. The A layer (at height $0$) consists mostly of pentagons linked by corners; the B layer (at $c/2$) also consists of pentagons, but these are bigger and share edges instead. The basic motifs in the A/B layers are therefore pentagonal antiprisms. Both the A and B layers contain a mix of L and S atoms. Between the A and B layers at $c/4$ and $3c/4$ are the much sparser X and $\rm X'$ layers. These contain only S atoms centering the pentagons in the A/B layers, thus forming chains of interpenetrating icosahedra along the $c$ direction. The X and $\rm X'$ layers are identical in projection on the quasiperiodic plane. The complete stacking sequence is $\rm AXBX'$. In order to discuss the structure in greater detail, it is convenient to present an idealized model; we shall do this in the next section. The centers of the columns of icosahedra are shown linked by added lines in Fig.~\ref{fig-proj}. Observe that the structures we found in simulations are not strictly periodic in the stacking direction: the axis centering a column of icosahedra may jump in position from one layer to the next. We also observed icosahedral structures under the same simulation conditions. Consequently, both phases might coexist in single (larger) sample. Such a coexistence is well-known experimentally in the Al-transition metal structure family, most commonly among metastable, rapidly-quenched quasicrystals but also for stable phase i(AlPdMn) and d(AlPdMn) \cite{alpdmn}. The simultaneous stability (or near stability) of both structures suggests that the local order is very similar. Indeed, we shall show that our decagonal structure model is quite closely related to the well-known `Henley-Elser' \cite{henels} model. \section { Decoration Models} \label{dsdm} The most convenient framework for making sense of our structures, without any bias about the kind of long-range order, is a decoration of a tiling with atoms. The same decoration scheme, with a finite number of site types and positional parameters, can model a random structure as a decoration of a random tiling, or an ideal quasiperiodic structure as a decoration of a quasiperiodic tiling. However, in attempting to describe an idealized structure model, we suffer from an embarrassment of riches. That is, there are several related structure models describing this kind of order. The models differ by (i) breaking a local symmetry in the position of an atom (which introduces matching-rule-like arrows) (ii) by allowing additional kinds of tiles (iii) by constraining the packing of tiles (this may create larger tiles as composites of smaller ones). Structure models are conveniently presented using idealized positions, represented as integer combinations of some finite set of basis vectors, but obviously the real atoms undergo displacements from these positions. Two structural models are called ``topologically equivalent'' (Mihalcovi{\v c} et al.\ 1996a) \nocite{mihI} when they converge to the same structure after equilibration at low temperature (or at zero temperature by relaxation under the pair potentials. Although apparently different, the models are physically indistinguishable. Thus the concept of topological equivalence is crucial in comparing different structure models and in organizing the above-mentioned family of related structure models. We will begin (Sec. \ref{basictiles}) with a simple simple tiling of three tiles -- the fat and skinny Penrose rhombus plus distorted hexagon ``Q'' -- from which the other tilings can be derived. This is not itself a good tiling model, because the packing is poor in the skinny rhombus, but it functions as a ``zero-order'' approximation for constructing better models. Therefore, we go on to consider two different modifications which correct this: (i) the triangle-rectangle tiling (Sec.~\ref{trirect}). and (ii) the ``two-level tiling'' (Sec. \ref{twolevel}) The relation of each modified model to the Penrose tiling decoration is an example of a ``differentiation'' in the terminology of Mihalkovi\v{c} et al.\ (1996b); \nocite{mihII} this means that a single environment in one model corresponds to several subclasses in the second model, which have somewhat different atomic decorations. \subsection{ Basic Tiles}\label{basictiles}\label{basic} Next we present the simplest possible version of the model, phrased as a decoration of the Penrose rhombi without matching rules. \subsubsection{ Penrose Rhombi}\label{penrose} Consider the decoration of the fat and skinny Penrose rhombi, as shown in Fig.~\ref{fig-basic}(a,b). This structure is not the best packed, but it is certainly one of the simplest. Each large atom is at a position dividing the long diagonal of a fat rhombus in the ratio $\tau^{-1}:\tau^{-2}$ where $\tau$ is the golden mean $(1+\sqrt 5)/2$. Notice that there is a complete symmetry between the A and B layers. Furthermore, the decorating atoms sit on the midpoint of each tile edge, so the decoration does not enforce the Penrose matching rules: it may be applied to any random tiling of these rhombi. The centering atoms of the pentagons are called $\alpha$. They are located in the X layer at the corners of the rhombi. In the A layer, the mid-edge atoms are called $\beta$ and the rhombus interior atoms $\lambda$. In the B layer, the equivalent mid-edge atoms are called $\gamma$ and the rhombus interior atoms are $\mu$. There are two alternative criteria for assigning the species: either according to interatomic distances in the model, or according to coordination number \cite{henels}; the two criteria are obviously correlated and (for all models presented in this paper) they lead to the same assignments. We assign $\beta/\gamma$ mid-edge sites to be S atom, since they have coordination 12 and tend to be squeezed by the $\alpha$ atoms at either end of the edge. The $\alpha$ sites are also assigned S atoms, since they have coordination 12 and are squeezed along the $z$ axis where they form chains. (The interlayer spacing is less than an interatomic spacing, since the other atoms don't form vertical chains.) Finally, the $\lambda/\mu$ sites, which all gave coordination 16, are taken to be L atoms. \subsubsection{ Q Tile } In the Penrose decoration, described so far, the A and B layers are equivalent by a $10_5$ screw symmetry, and correspondingly $\beta$ is equivalent to $\gamma$ and $\lambda$ is equivalent to $\mu$. However, in the decorations we present later, the A and B layers will be inequivalent. In those more complicated structures, a single Greek label refers to several subclasses of atoms having similar but not identical local environments. (Usually their coordination shells are topologically identical, but there is some variation in which neighbors are L and S atoms.) The Penrose tiling structure forces a ``short'' linkage between two vertices when they are related by the short diagonal of the skinny rhombus. Here ``linkage'' denotes a connection between the centers of neighboring motifs, which form the vertices of a rigid network. (In this case, the motif is the chain of centered pentagons, and the network is the tiling.) The $\alpha$ atoms projecting to those two vertices have {\it non}-icosahedral coordination shells. (There are similar to the awkward coordination shells occurring in the oblate rhombohedron \cite{henels}). We believe these are unfavorable energetically. (Indeed, nothing resembling a skinny rhombus is found in Fig.~1.) Therefore, we shall make an additional hypothesis: {\it true chains of icosahedra are energetically favorable and so their frequency should be maximized}. To increase the number of chains of icosahedra, we must somehow eliminate the short linkages. This is possible whenever two skinny tiles and a fat tile form a distorted hexagon; they can be joined in a composite tile Q as shown in Fig.~\ref{fig-basic}(c). The only change in atomic positions in forming the Q tile is along the axis over the central vertex, where the the $\alpha$ atoms forming the chain are replaced by $\lambda'$ atoms (in A layers, if the adjacent atoms along the tile edges are in a B layer, and otherwise vice versa). Note there are half as many new atoms $\lambda'$ as there were of old $\alpha$ atoms; they should be L atoms because of this larger spacing (or alternatively, because they have coordination 15). We will call this operation a ``chain-shift''. Note that the Q tile is considered to have a symmetry between the top apex and bottom apex (the light dashed lines in Fig.~\ref{fig-basic}(c) only illustrate the topological equivalence to the apparently asymmetric Penrose tile decoration.) The Q decoration should have the full symmetry of the Q tile's exterior; after the chain-shift, this require only small adjustments of position and species. The $\lambda$ site from the fat tile is renamed $\lambda'$ since it has coordination 15 after the chain-shift and is symmetry-equivalent to the $\lambda'$ site from the internal vertex. The non-Frank-Kasper $\gamma$ atom on the edge shared by skinny rhombi is converted into a $\mu$ L atom (upper part of Fig.~\ref{fig-basic}(c)) just like the $\mu$ site from the fat rhombus. Finally, the other two non-Frank-Kasper $\gamma$ sites on internal edges are renamed ``$\delta$'' and must be shifted slightly so as to lie symmetrically on the bisector of the long Q diagonal (since their neighbors are symmetric around that bisector.) They divide the bisector line almost into thirds (we used $0.312:0.376:0.312$ in the initial configuration for simulations in Sec.~\ref{secopt}). Since the $\delta$ atoms have coordination 14, and are somewhat squeezed with their neighbors along the bisector line, they have been designated S atoms. (However, note that coordination $14$ was found to be borderline in the closely related icosahedral structure: analogous sites occurring on the short axis of the rhombic dodecahedron tile (see below) are occupied by small atoms in i(AlCuLi) but by large atoms in i(AlMgZn)\cite{henels}). To maximize the number of proper centered-pentagon chains, then, the original tiling should be arranged with skinny rhombi grouped in pairs wherever possible. This is essentially the same mathematical problem as that of packing equal disks on Penrose tile vertices \cite{hen86}; those the disks correspond to the pentagon-chain motif in the present model. \subsubsection{ Icosahedral Tiling Relationship} \label{icorel} It is interesting to note the close relationship of this tiling to the three-dimensional ``Henley-Elser'' decoration of the Ammann rhombohedra by L and S atoms as a model of Frank-Kasper icosahedral quasicrystals such as $i$(AlCuLi) and $i$(AlZnMg). Let us orient rhombohedra so that one edge lies in the vertical direction as shown in Fig.~\ref{fig-PROR}, but shear them slightly so that the vertical edge has length $c$ and the other five kinds of edge have vertical component $c/2$ and horizontal components of length $a$ along one of the five basis vectors. Projecting the prolate and oblate rhombohedra onto the plane perpendicular to the selected axis generates respectively the fat and skinny rhombus of the (2D) Penrose tiling. Thus, every tiling of the rhombi can be extended vertically as a stacking of rhombohedra with period one edge length along the special axis. The layer of this stacking is a puckered slab dual to a grid-plane, analogous to one of the ``tracks'' in a two-dimensional Penrose tiling \cite{dunkat}. Note that, in contrast to a generic rhombohedral tiling, {\it every} rhombohedron in this stacking has edges in the vertical direction. In fact, it turns out that our decagonal decoration of Penrose tiles is ``topologically equivalent'' to the icosahedral decoration of rhombohedra by Henley \& Elser (1986) \nocite{henels} \footnote {A similar, but quite imperfect, relationship can be made between the icosahedral $i$(AlMn) (Elser \& Henley 1985; Mihalcovi{\v c} 1996b) \nocite{elshen,mihII} and the decagonal model based on the Al$_{13}$Fe$_4$ structure \cite{hendec}}. To see this, cut the rhombohedra along horizontal planes such as those shown dashed in Fig. 2(a,b) -- all the atoms in the PR and OR lie in or near these rather densely packed layers. Then reassemble them into prisms with horizontal rhombic faces. The atom positions form a decoration of the prisms, shown in Fig.~\ref{fig-3D}(c,d), which is clearly equivalent to that in Fig.~\ref{fig-basic}(a,b). (The ideal positions by Henley \& Elser (1986), \nocite{henels} projected onto the plane, are {\it exactly} those used here in the tiling; the vertical positions of the $\lambda$ atoms, however, must be adjusted by $\sim 0.026 c$.) The third tile of the Henley-Elser model, the rhombic dodecahedron (RD), is a little bit more complicated. (see Fig.~\ref{fig-RD}). Before it can be stacked, two skew triangular prisms at opposite sides of the RD, forming together one PR, must be removed. If this truncated RD is cut into layers and reassembled, it produces an elongated hexagonal prism which is decorated exactly as the Q tile in Fig.~\ref{fig-basic}(c). The structure obtained by transforming the Henley-Elser structure not only has the atoms in exactly the same positions (after the small vertical shifts) as our decagonal model, but also has exactly the same assignment of S and L atoms. That is not very surprising, since the species were assigned according to the local environments in both models. \subsection{ Triangle-Rectangle Tiling} \label{trirect} Returning to the 2D tiling, there are now two possible schemes to handling the skinny rhombi everywhere in the tiling. In one scheme, we imagine first that {\it all} skinny tiles are paired and absorbed into Q tiles: a generic tiling of this sort is the ``fR/Q (fat rhombus and Q) tiling''. Then the Q tile can be divided into a rectangle (R) and two isosceles triangles (T), with new edges of length $b\equiv (1+\tau^{-2})^{1/2} a$, while each rhombus can be divided into two isosceles triangles of the same kind. The two objects can form random tilings of the plane, most of which cannot be regrouped into rhombi \cite{ox}. The decoration introduced in this paper is compatible with an arbitrary triangle-rectangle (T/R) tiling. Note that, in this variant of the structure model, each tile comes in an A and a B flavor, denoting the level at which its mid-edge atoms sit; tiles adjoining each other by a $b$ edge have opposite flavors. The triangle and rectangle have long been known as the building blocks for the so-called pentagonal Frank-Kasper phases (see for example Samson (1968); Samson (1969)). \nocite{sam1,sam2} The 12-fold symmetric quasicrystals, which are typical Frank-Kasper phases, are built from a triangle and square on which the atomic arrangements are similar, but not identical, to those in the T and R tiles in our model. Furthermore, a T/R description like ours was used by He et al.\ \citeyear{chinese} to describe the $\rm Fe_{52}Nb_{48}$ quasicrystal, the only real material claimed to be a Frank-Kasper decagonal. (See Sec.~\ref{comparison}; Nissen \& Beeli (1993) \nocite{beeli} also used a rhombus-triangle-rectangle tiling to analyze TEM images of a Fe-Nb decagonal quasicrystal.) However, the decagonal T/R structure has been systematically studied as an abstract (random) tiling only recently \cite {ox,cock}. We shall now consider energy energy differences between different T/R tilings which might cause additional ordering. It would seem unfavorable energetically for two rectangles to adjoin along an $a$ edge, since the $\delta$ and $\beta/\gamma$ atoms along the bisecting axis are tight within each tile and would be squeezed against each other; let us add this to the constraints defining our tiling ensemble. Then Cockayne's quasiperiodic T/R packing\cite{cock} satisfies the constraint just mentioned; it satisfied a second constraint, in that rectangles don't adjoin along $b$ edges, either. With the latter constraint, every rectangle must be capped with triangles forming a complete Q tile, and then all other triangles must be paired along their $b$ edges forming complete fat rhombi: thus the structure is more economically described as an fR/Q tiling with the constraint that Q's may not lie side-to-side. (We presume the ensemble of such packings is a random tiling, but this is not proven.) Every fR/Q structure, of course, can be considered as a fat and skinny (fR/sR) rhombus tiling produced by decomposing the Q tile. It has been pointed out \cite{mihox,ox} that an fR/Q tiling gives the optimal disk packing (i.e. maximal density of $\alpha$-chains, in our models) of all Penrose tiling, since {\it all} the skinny rhombi are grouped into pairs, unlike the two-level tiling described below. Possibly the additional constraint on the T/R tiling is justified by the energetic cost of adjoining along the $b$ edge -- the $\lambda'$ atoms inside one R press against the $\delta$ atoms (in the same layer) inside the other R -- but we have not identified any compelling argument and hence do not propose the fR/Q tiling as the optimal geometry. \subsection{Two-Level Tiling} \label{twolevel} The decagonal structure observed after cooling from the melt (Sec.~I) had the additional property that its A layer is inequivalent to its B layer. We shall now turn to a structure model which adopts this as a postulate. Chain-shifts occur only at vertices where all the edges are decorated with B layer atoms. Naturally, the B atoms in turn shift away from the midpoint of the edge towards the partially vacated columns. (This was observed in the simulation, in the fact that the B layer pentagons were bigger than A layer pentagons.) This has the effect of the double-arrow matching rule in the Penrose tiling. An arbitrary tiling obeying only the double-arrow matching rules (Tang and Jari\'c, 1990) is equivalent to a random tiling of three composite supertiles Q, K, and S\footnote{ The three tiles are called ``Q, K, S'' after the corresponding three local environments found in the Penrose tiling where double-arrow edges can meet. The same tiles are called ``h,c,s'' respectively by Li \& Kuo (1991). \nocite{li2} } which are obtained by merging Penrose tiles along the double-arrow edges. Since the atoms decorating the outer edges of the Q, K, and S tiles remain symmetric about the middle of the edge, they do {\it not} enforce the (single-arrow) Penrose matching rules on these edges: all random tilings of the plane by Q, K, S tiles are a priori permissible. The random tiling of these tiles has been called the ``two-level'' tiling (2LT)\cite{henART}; these tiles were independently proposed by Li \& Kuo (1991). \nocite{li2} Fig.~\ref{fig-2LTdec} shows a fragment of the 2LT, including one of each tile. As a special case, the Penrose tiling, modified by combining rhombi along their double-arrow edges, becomes a simple quasiperiodic Q, K, S tiling. The ``Q'' tile decoration was described already; the star tile S is also straightforward, being a composite of five fat rhombi. Note that no chain-shift is necessary at the center of the S tile since there is no skinny tile adjoining; thus this site forms an exception, around which the A/B symmetry breaking (between large and small pentagons) is opposite to the pattern around all other $\alpha$-chains. Otherwise, all the edge sharing pentagons are in layer A and the corner sharing pentagons are in layer B. The ``K'' tile is a grouping of a skinny rhombus with three fat rhombi; there is a chain-shift on the central vertex, which becomes known as a $\kappa$ site. This is the second option for absorbing the extra single skinny rhombi, instead of pairing all of them in Q tiles. However, the atomic arrangements in the K are still somewhat awkward. The four $\gamma$ atoms surrounding the internal vertex can only relax to form a sort of pentagon with a missing corner, meaning that the $\kappa$ coordination shell is not a Frank-Kasper polyhedron but a distorted octahedron. We could have arrived directly at the two-level tiling from analysis of Fig.~\ref{fig-proj}, if we accepted as a postulate that in the linkage between two chain-motifs, the edge-sharing pentagons are always in the B layer. We quickly arrive at a geometry where the angles between linkages are all multiples of $2\pi/5$. The smallest tiles enclosed by such edges are the Q, K and S tiles of the ``two-level'' tiling (Fig. \ref{fig-2LTdec}). The details of this derivation are in Appendix A. Reviewing Sec.~\ref{dsdm}, we can identify a hierarchy of successively more ordered tiling geometries. The Penrose tiling is the most ordered; it can be broken up into the ``two-level'' tiling of Q, K, S tiles. The K and S tiles can be broken up further into fat and skinny rhombi; finally the fat rhombi can be broken into triangles, and the Q tiles broken into triangles and rectangles. \section{ Quasiperiodic Models and Stoichiometry} \label{environments} Even when a quasicrystal model has been formulated as a tile decoration, it is often illuminating to represent it as a quasiperiodic structure obtained by a cut through a higher-dimensional space, since (i) the space-group and diffraction properties are most concretely understood in this fashion (ii) different site types are represented as domains of the acceptance domains in the ``perp'' space, thus the relations between different site types are visible in the spatial relation of the domains. (iii) The number density (or frequency) of a particular species or site type is proportional to the area of its portion of the acceptance domain, which is convenient for computing a definite stoichiometry. This description is helpful even for atomic models based on random tilings. These, when ensemble-averaged, produce structures of partially-occupied sites. Mapped into perp space, such distributions can usually be {\it approximated} as quasiperiodic acceptance domains convolved with a Gaussian in perp space\cite{henART}. The first step is to find a quasiperiodic tiling. This is trivial for the two-level (Q/K/S) tiling, but highly nontrivial for the T/R (or for the fR/Q) tiling; in that case, the tiling can only be constructed by an elaborate deflation, and the corresponding acceptance domains have fractal boundaries \cite{cock}. Therefore, we will limit ourselves to presenting the acceptance domains based on the Q/K/S-tiling decoration, where the tiles are arranged in the perfect Penrose tiling. (A Penrose tiling is turned into a two-level (Q/K/S) tiling simply by erasing the double-arrow edges). Also in this section, we tabulate the contents in atoms of each tile. This approach to stoichiometry is a more powerful than the approach using the acceptance domain, since it can be applied even to random tilings.\\[1ex] \subsubsection{ Idealized Structure and Acceptance Domain} In order to construct the acceptance domain, we first make a crude idealization of the Penrose-rhombus decoration which we presented first (Fig.~\ref{fig-basic}(a,b)). All $\beta$ atoms lie on single-arrow edges and all $\gamma$ atoms lie on double-arrow edges. Let us shift the $\beta$ atoms and the $\gamma$ atoms in the direction of Penrose's arrows, such that each divides its edge in the ratio $\tau^{-1}:\tau^{-2}$. (This ratio is technically convenient in allowing the ideal coordinates to be written as integer combinations of the five basis vectors. They are actually most simply treated using a deflated Penrose tiling basis with lattice constant $\tau^{-1} a$.) In going to the Q/K/S tiles for in the quasiperiodic structure, we retain these same positions but change the label and the species as described in Sec.~\ref{twolevel}; note in particular that the displacement properly places the $\delta$ atoms on the bisecting symmetry axis of the Q tile. The B layer forms Penrose's packing of regular pentagons, one centered on every vertex of the 2LT. Each ``internal'' vertex (where double-arrows meet) is also surrounded by a pentagon, but these are irregular and have many short distances. The acceptance domains are shown in Fig.~\ref{fig-acc}. Note there is a separate domain for layers A, B, and X of the real structure; in addition, as always, the vertices project into five flat two-dimensional layers in three-dimensional perp-space, so we must show separate shapes for each perp-space layer. Finally we have shaded the domain with dark and light shading to distinguish L and S atom occupancy. Although this structure has unreasonable distances, it is ``topologically equivalent'' (Mihalcovi{\v c} et al.\ 1996a)\nocite{mihI} to the (reasonable) version presented in figure \ref{qksdecl}. We could alternatively compute the acceptance domains for the decoration in figure \ref{qksdecl}, with the atoms moved to mid-edge positions (except for those in the Q tile). Relative to Fig.~\ref{fig-acc}, the $\beta$, $\gamma$, and $\delta$ subdomains each would get broken up into five pieces, and shifted in the parallel space direction. Since the mid-edge atom positions are also special crystallographic positions, several of the shifted subdomains will reunite there and form a new piece of acceptance domain which is centered on a mid-edge site in the 5-dimensional hypercubic lattice. We forgo to present these domains since this second picture is less compact and not as appealing as the first one. The space group of this quasiperiodic decoration is $P10 /m m m$. In other variant decorations in which the A/B symmetry is not broken, we would get an additional system of mirror planes perpendicular to the 10-fold axis and this changes the space group to $ P10_{5} /m m c$. The point group of the motif with the highest symmetry is $5 /m m$ \cite{hendec}. \subsubsection{ Stoichiometry} The most general approach to stoichiometry is to count the number of atoms associated with each tile. Table I summarizes the contents of each tile. The combination Q+S has contents 29L+14S while 2K (which covers the same area as Q+S) has contents 28L+14S. This suggests that the K tile is too loosely packed; indeed, the average coordination number in K is also a bit low. Table II gives the coordination numbers and the net stoichiometry for each decoration. In the case of the 2LT (Q/K/S) decoration, this agrees with the stoichiometry which can be read off from Fig.~\ref{fig-acc}. It can be seen that the stoichiometry is close to $L S_2$ in both cases. (The pure Penrose tiling decoration, an unrealistic model, has $L_{0.198} S_{0.802}$.) That seems to be considerably fewer L atoms than in the T(AlZnMg) icosahedral approximant \cite{berg} or Samson's AlMg-type structures, but it is comparable to i(Al$_6$CuMg$_3$) in which only $30\%$ of the atoms are L (Mg) \cite{sam1,sam2}. \section{ Further Simulations} \label{XXX} In section III, we worked out the geometric properties of the idealized model, in several closely related variants based on different tiling rules. We derived the coordination configuration for all sites: all atoms have a reasonable distance from their neighbours and most of them are tetrahedrally close packed. The idealized model, in any of the variants presented in the last two sections, is well-packed and (nearly) tetrahedrally close-packed around every atom. Hence it is expected to be (locally) stable against disordering by thermal fluctuations. To check this stability, we ran additional simulations starting with a configuration from the ideal model. A major motivation of these simulations was to locate the region of parameter space in which the quasicrystal phase may be thermodynamically stable; hence, we attempted to adapt the potentials to the structure model, as explained in Subsec.~\ref{secopt}. As a result of the optimization there we can predict the ideal c/a ratio and the density of the structure, as well as the ratios of bond radii for LL, LS and SS pairs which seem most conducive for (decagonal) quasicrystal-forming. We used not only Lennard-Jones (LJ) potentials as in eq.~(\ref{eq-LJ}) but also the Dzugutov potential, with independent parameters for the three ways of pairing species $\alpha$ and $\beta$. Both pair potentials have a single attractive well with a minimum at radius $2^{1/6} \sigma_{\alpha\beta}$, and having depth $\epsilon_{\alpha\beta}$; here $\sigma_{\alpha\beta}$ and $\epsilon_{\alpha\beta}$ are called the ``bond parameter'' and ``interaction parameter''. The Dzugutov potential \cite{dzug} has an additional repulsive maximum at a radius $\sim 1.6$ times that of the minimum, and a height $\sim 0.5 \epsilon _{\alpha\beta}$; this is designed to disfavor the square arrangements found in fcc, bcc or hcp structures. \subsection{ Optimization of Parameters} \label{secopt} To find the optimal potential parameters, we minimized the potential energy at $T=0$ while varying these parameters. For this purpose we used a fixed ideal structure model with the ideal atom position and without relaxation. The sample size was a $p/q~=~3/2$ approximant\footnote{ the basis vectors are $(2 q, 0), (p-q, p), (-p, q), (-p, -q), (p-q, -p)$, and the periods are\\ $2(2 p-q) \tau+2(3 q-p)$ in $x$-direction and $4 \sin(\pi/10)(p \tau+q)$ in $y$-direction.} of a perfect Penrose tiling with 10 periods of XAXB layers and ideal size $l_x \cdot l_y \cdot l_z = 18.95 \cdot 16.09 \cdot 22.93$, containing 3270 small and 1420 large atoms. Samples of different size do not yield different results within the accuracy we achieve. In a three-dimensional structure with two species of atoms L and S, bond and interaction parameters $\sigma_{\alpha\beta}$ and $\epsilon_{\alpha\beta}$ must each be determined for pairs $\alpha\beta$ equal to $LL$, $LS$, and $SS$. Of course, one of the $\epsilon_{\alpha\beta}$'s can be eliminated in principle, in the choice of the energy unit. We chose the sample size/shape (which was kept fixed in the $xy$ plane of the layers) so as to the Penrose tile edge length at exactly $a=2$ (the nearest-neighbour distance is roughly $a/2$.) Thus all three $\sigma_{\alpha\beta}$'s are nontrivial parameters. Results obtained in optimizing potentials for binary icosahedral Frank-Kasper structures (Roth et al.\ 1990)\nocite{rotstab} with similar local structure and potentials, show that $\sigma_{\alpha\beta}$ are largely independent of $\epsilon_{\alpha\beta}$, for $\epsilon_{\alpha\beta}$ in the range of $0.5 < \epsilon_{\alpha\beta} < 2.0$. Therefore we set $\epsilon_{LS} \equiv 1$ and $\epsilon_{LL} = \epsilon_{LS} \equiv 1/2$. Thus, we actually varied only the three $\sigma_{\alpha\beta}$'s. Before we can carry out the optimization itself we have to determine the lattice constant $c$ in the stacking direction. (The A/B layer spacing $c/2$ should also be roughly a lattice constant.) This is done in a poor man's minimization by scanning the interval $0.8 < c < 4.0$ and looking for the minimum of the potential energy $E(c)$ in the following way: having fixed $a$ (and consequently $l_x, l_y$) we calculate the partial radial distribution functions $g_{\alpha\beta}(r)$ for the ideal model using a given layer distance $c$. Then, without annealing or relaxing, we compute the partial potential energy \begin{equation} E_{\alpha\beta}(\sigma_{\alpha\beta}, c) = \int_{0}^{r_{c}} v_{\alpha\beta} \left(\frac{r}{\sigma_{\alpha\beta}}\right)\, g_{\alpha\beta}(r, c)\, 4\pi r^{2}\, dr \end{equation} dependent on the cut-off radius $r_{c}$ and the pair potential $v_{\alpha\beta}$ and minimize it as a function of $\sigma_{\alpha\beta}$. Using this procedure we find that the total potential energy $E = E_{LL} + E_{LS} + E_{SS}$ is minimzed if each $E_{\alpha\beta}$ is optimized separately. The optimal $\sigma_{\alpha\beta}$ for Lennard-Jones potentials are $\sigma_{SS} = 1.029$, $\sigma_{SL} = 1.137$, and $\sigma_{LL} = 1.189$. For the potentials used by Dzugutov we get $\sigma_{SS} = 1.108$, $\sigma_{SL} = 1.034$, and $\sigma_{LL} = 0.929$. The result reflects the fact that the Dzugutov potentials are very short ranged and that they have a maximum repulsive for second nearest neighbours, whereas even in the cut-off and smoothed version of the Lennard-Jones potential interactions with the second or third set of neighbours are still attractive. We also found that the optimal A/B interlayer distance for Lennard-Jones and for Dzugutov potentials is about $c/2= 1.04\pm 0.01$ The uncertainty quoted here reflects the slight variation of the spacing depending on the choice of $\epsilon_{\alpha\beta}$ and $r_{c}$. \subsection{ Results} The constant temperature and pressure molecular dynamics simulation method described in Sec.~\ref{simulations} has been applied to study the thermodynamic stability of the ideal structure. To simulate non-cubic structures it has been extended to allow independent changes of the box lengths $l_{k}$. Thus during the MD simulations the box size could vary in contrast to the optimization, were the box was fixed. The simulation sample was the same as the one used in Sec.~\ref{secopt}. With Dzugutov's potentials the melting temperature at $P^{*} = 0.01 $ is $T^{*}_{\rm m} = 1.23$. (In fact, the structure does not melt; it vaporizes.) Of course, in an infinite box the ratio $l_x/l_y = 2 \sin 36^\circ \sim 1.1756$ is fixed by pentagonal symmetry. In a finite system, however, $l_x/l_y$ should deviate slightly from the ideal ratio due to the ``phonon-phason'' coupling (Lubensky, 1988, and references therein) since the periodic boundary conditions force a nonzero phason strain. We found the equilibrium value $l_y/l_x = 1.177$ in the simulation, independent of the temperature and very close to the ideal value. The A and B layers are mirror planes so they should still be flat in the relaxed structure; on the other hand, the X layers should pucker slightly after relaxation, with the $\rm X'$ layer puckered in the opposite direction. The simulation, however, show that the layers remain essentially flat. We have also recorded the atomic mobility by monitoring the mean square displacement of the atoms. It turns out that even close to the melting point ($T = 0.9 T^{*}_{\rm m}$) long-range diffusion is {\em not} observable at all, and only very few atoms are seen to jump to a new position. Those atoms which do jump are all in the skinny rhombus part of the K-tile. Similar results have been obtained for icosahedral and dodecagonal structures \cite{rothdiff} where oblate rhombohedra and skinny rhombic prisms resp.\ play the same role as the flat rhombus in the K-tile. Most of the jumps in the K tile take place around the $\alpha$ atom at the top of the tile (see Fig. \ref{qksdecl}) and around the interior vertex. Possibly the jumping in this environment indicates that our model is erroneous there: perhaps a different arrangement or choice species should occur in the atoms surrounding the interior vertex of the K tile, or perhaps the correct model is a T/R tiling which has no K tiles at all. It is also possible that the correct structure model should be thought of simply as a random tiling of Penrose rhombi (Sec. \ref{penrose}); then ``phason fluctuations'' are realized by tile flips that entail switching which vertex of a skinny rhombus undergoes a ``chain-shift''. Our conclusion is that the structure is very stable, and that the choice of the atoms positions is reasonable. The jumping atoms do not affect the stability since they are separated by large distances. \section{ Discussion} \subsection{ Summary} Now we are back at the beginning: we started with a structure found by computer simulations, and recognized that the structure could be described by a tiling model. We derived quite a number of tilings envolving fR, sR, T, R, Q, K, and S tiles, all of which are compatible with a decoration with atoms that include the original model. Although the fR/Q and T/R models seem to be superior to the Q/K/S tiling and the variants involving sR tiles, there may be properties that favour the latter. We tested one variant (Q/K/S) of the tiling models again with computer simulation to derive structure and potential parameters, and to find out if the geometically reasonable model is also reasonable with a simple interaction model. Since the Q/K/S tiling is somehow cumbersome, but on the other hand includes most of the properties of the other tilings, its stability assures us that the other tiling types will also be stable. Up to now we have only treated each tiling for itself and described its properties. In this final section we would like to address the interrelationship of the tiling types, and if there is a hierarchy of tilings. This includes the A/B layer symmetry breaking which occurs in some of the models. In addition to computer simulations it will be interesting to compare the tiling models to experiment and to other structure models. Furthermore we want to ask if and how it is possible to change a certain tiling which means reshuffling the tiles, introducing defects like random stacking and atomic jumps. These moves may also lead to the a transformation of one tiling type into another. Although such a process is not suitable for MD simulations, it may be studied in Monte Carlo simulations with properly chosen interactions. \subsection{ Hierarchy of Tiling Descriptions} Our approach is an example of a general approach which may be fruitful in understanding the relation of interatomic interactions to long-range order in quasicrystals. An ensemble of tilings is proposed; via a decoration rule, these correspond one-to-one to an ensemble of low-energy atomic structures. However, the energy is slightly different for each of these structures; these energy differences may be considered as a ``tiling Hamiltonian'' ${\cal H}_{tile}$ which is purely a function of the tile arrangement. Often the tiling Hamiltonian is a sum of one-tile and neighbor-tile pair interactions (Mihalcovi{\v c} et al.\ 1996a,1996b) \nocite{mihI,mihII}; alternatively, Jeong and Steinhardt (1994) proposed a family of tiling Hamiltonians with a (favorable) energy $-V$ for each occurrence of a special local pattern, which they called the ``cluster''. The ground states of ${\cal H}_{tile}$ are a subset of the original ensemble; most often this constrained sub-ensemble can be described by a tiling of larger (``super'') tiles with their own packing rules. In the ``cluster'' Hamiltonians studied by Jeong and Steinhardt (1994), the degree of long-range order is enhanced in the supertiling -- indeed for one special ``cluster'' one just obtains the quasiperiodic Penrose tiling -- and even a random tiling of supertiles has greatly reduced phason fluctuations. Evidently, if ${\cal H}_{tile}$ can be divided into a sum of successively weaker parts, there may be a corresponding hierarchy of successively larger supertiles. Furthermore, as the temperature is raised, the constraints due to the weakest terms of ${\cal H}_{tile}$ are broken, so the size of the relevant tiles may decrease with temperature. (The associated changes in diffraction pattern were discussed by Lan\c{c}on, Billard, Burkov \& de Boissieu (1994)). \nocite{burkov} In Sec.~\ref{dsdm} , we discovered just such a hierarchy of similar structures, some more constrained than others, as follows: the fR/Q tiling is a special case of either the fR/sR/Q (fat and skinny rhombus and Q) tiling, or of the T/R tiling. The 2-level (Q/K/S) tiling is also a special case of the fR/sR/Q, which is a modification of the fR/sR tiling (Subsec.~\ref{basic}). We postulated (at the start of Sec.~\ref{dsdm}) that the dominant term in our tiling-Hamiltonian favors the chain-motif, consisting of $\alpha$ atoms surrounded by alternating pentagons. (Incidentally, this is a ``cluster'' in the sense of Jeong and Steinhardt (1994).) Then the sub-ensemble which minimizes that term is the T/R random tiling, which (see below) is also consistent with a structure model proposed by He et al.\ \citeyear{chinese}. Due to the limitations of our simulation, and because we did not perform relaxation at $T=0$ to extract the parameters of ${\cal H}_{tile}$, we could not prove that the chain-motif (and hence T/R tiling) is favored, nor could we measure the smaller energy differences that would decide which subensemble of T/R tilings most faithfully models the structure. We could only conjecture (in Sec.~\ref{trirect}) that there should be an additional term in ${\cal H}_{tile}$, disfavoring Q tiles adjoining by $a$ edges. Granting that, it followed that the best tiling to represent our structures is something between the T/R tiling and the fR/Q tiling of Cockayne (1994). \nocite{cock} \subsection{ Issues of Symmetry Breaking in our Simulation} There remains a serious doubt that the apparent decagonal structure we observed in our simulations might be only an artifact of the geometric restrictions imposed by the small simulation box: the correct phase in the thermodynamic limit might be an icosahedral random tiling. Indeed, Sec.~\ref{icorel}, showed that our decagonal structure is essentially a special layered packing of the icosahedral tiles; rhombohedral tiling might find a layered arrangement in a quench, if that happens to fit well with the periodic boundary conditions. Furthermore, out of seven quenches with the same potentials (Roth et al.\ 1995), \nocite{rotcool} five quenches led to icosahedral structures and two to decagonal ones. Even if we grant that the true structure {\it is} decagonal, there is a second question whether we may be over-interpreting accidental effects of the finite-size box used in our initial simulations: does an A/B symmetry-breaking occur in the thermodynamic limit? This is the fundamental difference between the 2LT (Q/K/S) model and the T/R tiling model. Apart from the observation in our initial simulation, there is no convincing reason for the 2LT (Q/K/S) model, since the K tile seems to be poorly packed. There is no other numerical evidence that the A/B symmetry-breaking really occurs in an extended structure. It is quite possible that the shape and size of each layer in our simiulation box happened to allow only a tiling which lacks the $180^\circ$ symmetry axis (out of the plane) which would make A and B edges equivalent.\footnote{ A simple matching rule has been given that favours the Q/K/S tiling over T/R and fR/Q: if you consider the pentagonal antiprisms around the $\alpha$ atoms, then there should always be two different atoms on the opposite sides of a diameter (true in the Q/K/S case). In the T/R tiling there are only two vertex environments that fullfil this condition, in the fR/Q tiling, there is only one. But with this vertex environments alone it is not posssible to produce a fR/Q or T/R tiling. Thus the Q/K/S tiling is favoured. The rule also works if you have sR tiles: In an sR/fR and an sR/fR/Q tiling it either forces the assembly of Q/K/S tiles or induces chain shifts. } To pursue this further, we should consider whether there is any physical cause for such a symmetry breaking. Rather than start with the Q/K/S tiles, it may be better to start by imagining a fR/sR/Q tiling and asking what might drive the symmetry breaking, which from this viewpoint is a displacive instability of the $\gamma$ atoms, perhaps coupled to some chain-shifts. (After all, on every skinny rhombus such a symmetry-breaking occurs locally; the question is whether the pattern of displacements can be propagated from one skinny to the next.) It is plausible that, once A/B symmetry breaking is granted, it will be advantageous to group the rhombi into S and K tiles. Both of these questions might be addressed (at zero temperature) by relaxing a variety of decagonal and icosahedral structures under a variety of potentials, to see which is more stable energetically, in the spirit of calculations done on some Al-transition metal quasicrystals (Phillips and Widom, 1994; Mihalcovi{\v c} et al.\ 1996b) \nocite{phil,mihII}. \subsection{ Comparison to Other Structure Models} \label{comparison} The Fe-Nb structure model outlined by He et al.\ \citeyear{chinese} was not formulated in terms of a tiling of any kind, nor as a deterministic rule for packing the plane with motifs; instead, it is largely a scheme for analyzing high-resolution images. Although there is a possibility that they misinterpreted images of a periodic approximant of the quasicrystal, or a microcrystalline mixture of approximants, it is still interesting and economical to decribe it using quasicrystal tiling element. Their model is based on the same motifs (pentagon-chains) as ours and the lines in figure 3(b) of He et al.\ \citeyear{chinese}, which indeed outline triangle and rectangle tiles, are the same as the $a$ linkages in our model. Since the sample they imaged was rather small and defective (much like out first simulation), we can only be tentative which variant of our structure model it should be identified with. In principle a T/R tiling could be a fR/Q tiling with the $b$ edges drawn in. However, the image of He et al.\ \citeyear{chinese} includes pairs of rectangles adjoining by a $b$ edge, which is a defect from the fR/Q viewpoint; note also it clearly does {\it not} have a A/B layer ordering. Thus it is most plausibly idealized as a random T/R tiling; The icosahedral relationship described in Sec.~\ref{icorel} allows the decagonal structure to grow on the icosahedral one. Congruent icosahedral/decagonal grain boundaries should therefore be possible. The decagonal phase could be viewed as an approximant to the icosahedral phase, and thin decagonal bands in the icosahedral phase may be regarded as stacking defects. Furthermore, since the icosahedral and the decagonal phases are similar in composition, phase transformations between them should also be possible. Thus our Lennard-Jones system might serve as a toy system for investigating the behavior $i$-AlPdMn. (However, we must re-emphasize that the atomic arrangement in an Al-transition metal quasicrystal certainly differs from the Frank-Kasper quasicrystal described here.) We turn briefly to another approach to constructing structure models -- that based on atomic clusters (Elser and Henley, 1985; Henley and Elser, 1986; Mihalcovi{\v c} et al.\ 1996a)\nocite{elshen,henels,mihI}, not to be confused with the ``clusters'' of Jeong and Steinhardt (1994)! Now, Al-Mn type structure models have a common motif of $\rm Al_6 Mn_4$ tetrahedra \cite{kreiner}; by joining several of these along their faces larger clusters are formed which are also observed in these structures. On the other hand, Frank-Kasper models have a common motif of ``truncated tetrahedra'' surrounding every L atom with coordination 16; indeed, this motif is frequent in our structures. (see Fig.~\ref{friauf}; the full coordination shell, with 4 additional atoms, is the ``Friauf polyhedron''.) These are combined into larger clusters \cite{sam1,sam2} in exactly the same fashion. Thus, if an Al-Mn model can be represented as packing of $\rm Al_6 Mn_4$ tetrahedra (plus atoms needed to fill the interstices), then by replacing each of these by a truncated tetrahedron we produce a (hypothetical) Frank-Kasper model. For example, the ``Mackay Icosahedron'' \cite{elshen} is a combination of 20 tetrahedra which maps to the Bergman polyhedron (Bergman et al.\ 1957)\nocite{berg}. It is well known that that crystal phases $\alpha$-AlMnSi and $R$-AlCuLi are bcc packings of the respective clusters, and this suggested that the same cluster-cluster networks could describe the related icosahedral quasicrystals i-AlMnSi and i-AlCuLi \cite{henCCT}. A combination of 5 tetrahedra produces a pentagonal-bipyramid motif which is the basis for crystalline approximants [such as $\rm Al_{13}Fe_4$ and $\rm T_3(AlMnZn)$] and for conjectured structure models of {\it decagonal} Al-transition metal quasicrystals \cite{hendec}. The corresponding combination in a Frank-Kasper structure [consisting of 5 truncated tetrahedra arranged around a central axis] is the ``VF'' cluster, a common motif in large-unit-cell Frank-Kasper alloys of simple metals \cite{sam1,sam2}. If we assume that such clusters are linked as in $\rm Al_{13}Fe_4$, we produce a new hypothetical Frank-Kasper decagonal model {\it different} from the one presented in the present paper. (The ``VF'' motif is rarer in our models -- it occurs only when five fat rhombi meet to form a star.) In the new structure model, the centers of neighboring clusters are separated by $\tau a$ in the horizontal planes and by $\pm c/2$ vertically. We have not observed or investigated such a model in simulations. \subsection{ Beyond Two Dimensions} To describe real structures, we must go beyond static, perfectly stacked structures. The real structure has 2 extra dimensions: time, and the periodic direction $z$. Whether its ultimate state is random-tiling or a locked quasiperiodic tiling, a tiling can improve its order under annealing only by reshuffling of its tiles. For example, recoognizing the fundamental reshuffling in the case of the binary tiling permitted an accelerated Monte Carlo move (Widom et al.\ 1987)\nocite{wid87}). Note that, although the tile rearrangement apparently involves a large volume, it is common that the reshuffling requires the motion of only a single atom \cite{wid92}. The problem of reshuffling has an obvious relation to that of stacking disorder: a realistic model must handle configurations in which the structure does not repeat precisely along the fivefold axis. Indeed, the structure as quenched was not strictly periodic in the stacking direction. Of course, such nonperiodicity may be considered a defect, and perhaps blamed on incomplete equilibration; however there are two senses in which it is necessary even in equilibrium. First, reaching equilibrium usually demands some tile reshufflings. But in a 3D atomic structure based on a 2D tiling, every tile reshuffling requires rearranging an entire column of the atomic structure. Obviously this is easiest done one bit of the column at a time; the intermediate state is one with stacking randomness. Second, in the random-tiling explanation of the thermodynamic stability of quasicrystals, the contribution of tile-reshuffling entropy to the free energy is decisive; this entropy can be extensive only in an ensemble with stacking disorder \cite{henART}. Recently, Ritsch, Nissen, and Beeli \citeyear{ritsch} have argued that common features in high-resolution transmission electron microscope images of d(AlCoNi) are evidence of stacking randomness in this decagonal.\\[1ex] \subsubsection{ Reshuffling}\label{reshuffling} The reshufflings in the case of the two-level tiling are simple and are shown in Fig.~\ref{fig-old2}. We can trade off $\rm Q+S \leftrightarrow 2 K$ (Fig.~\ref{fig-old2}(a)) or $\rm Q+K \leftrightarrow K+Q$ (Fig.~\ref{fig-old2}(b)), This reshuffling is the simplest fundamental move which can be used to visit from one state to another. Local reshufflings of the T/R tiling are not possible; instead it is necessary to form a ``defect tile'' such as a skinny rhombus that is normally absent in that tiling. A defect tile can move through the tiling in what is called a ``zipper move'' \cite {ox} since it leaves behind a rearranged structure. One must check what reshufflings imply for the atoms. In particular, the number of each species ought to be conserved; otherwise the reshuffling is either blocked, or it creates substitutional defects, or it only takes place in a structure which is already substitutionally disordered in equilibrium. In a rhombus tiling, the phason strain (which can be regarded as the deviation of the projection plane embedded in higher dimensional space from its ideal position) determines the number density of each kind of rhombus; since the rhombus content is conserved under reshuffling, so is the atomc content. The same thing is true in our T/R tiling, but not in the two-level (Q/K/S) tiling, because of the possibility of trading off $\rm Q+S \leftrightarrow 2 K$. (When divided into Penrose rhombi, both combinations have the same contents). Indeed, our 2LT (Q/K/S) atomic structure model has serious problems with the reshufflings: the $\rm Q+S \leftrightarrow 2 K$ reshuffling changes the content of the tiles from 29 L+ 14 S to 28 L +14 S, so one large atom disappears or has to become an interstitial. The $\rm Q+K \leftrightarrow K+Q$ reshuffling conserves the net atomic content of the tiles; however, the atomic rearrangement in which atoms jump the least distance would put atoms on the wrong site for their species. Simulations have shown that such substitutions are not acceptable; even a few of them destroy the stability of the structure. We have to admit that no better moves have been found up to now. In a T/R tiling, the question of rearrangements is quite different since the only possible update move is a ``zipper'', is an entire chain of tiles which closes on itself \cite{ox}, just as in triangle-square tilings \cite{oxhen}. Thus, the intermediate state of a rearrangement involves not only stacking defects, but also defects within each layer (which would appear as special tiles other than T or R). \subsubsection{ Stacking Randomness} In a nonperiodic stacking, we may presume the tilings describing adjacent layers are similar. Thus the spatial sequence of layers is much like the temporal sequence of a 2D tiling undergoing a series of reshufflings \cite{henART}. Each violation of periodicity presumably costs energy; most likely, one or two ways of doing so are less costly than any others, so it is a reasonable approximation to postulate as a {\it constraint } on the stacked tiling ensemble that layers can be related spatially only in such ways (point stacking defects). In the case of the Q/K/S tiling model, the obvious stacking defects are flips $\rm Q+S \leftrightarrow 2 K$ and $\rm Q+K \leftrightarrow K+Q$ from one layer to the next. Since the positions of the atoms are similar in both states, these defects should be not too costly in energy. An interesting experiment would be to create a stacking of layers from {\it completely different} 2D tilings, and see what structure it relaxes to. For the {\it dodecagonal} Frank-Kasper structure, in the monatomic case \cite{rothdiff}, the atoms rearrange -- while moving only short distances -- until all the layers are {\it identical} and defect free; on the other hand, in the binary case of the same dodecagonal structure (L and S atoms, similar to the present paper) the structure cannot be healed with just short-distance moves; instead substitutional defects are introduced. \section*{ Acknowledgments} J.R. would like to acknowledge a post doctoral fellowship from the DFG and the hospitality at Cornell University, Ithaca, where most of this work has been carried out. C.L.H. was supported by the U.S. D.O.E. grant DE-FG02ER89-45404.
proofpile-arXiv_065-488
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} \label{sec:intro} After ten years of research in the field of high-$T_c$ superconductors\cite{BedMue} (HTSC), many of their properties have not yet been understood. In particular, the symmetry of the superconducting gap\cite{Lyo,Dyn,Sch} is still controversial. Usually, one assumes that the superconducting condensate in the HTSC can be described by an order parameter $\Delta_{\mbf k}$, which depends only on the quasi momentum ${\mbf k}$, but not on band index $n$. Retardation effects are also often neglected, i.e.\ the gap is assumed to be independent of the frequency $\omega$. A wide range of experimental techniques can be employed to investigate the properties of the gap function. Among these, Raman scattering has played an important role.\cite{DevEin,KraCar} The dependence of the Raman response on the directions of polarization of the incident and scattered light yields several independent spectra which provide a considerable number of constraints on the assumed ${\mbf k}$-dependence of the gap function $\Delta_{\mbf k}$. However, Raman scattering is not sensitive to the phase of the gap. The Raman spectra at temperatures below $T_c$ shows, in most HTSC, a clear gap-like structure which lies in the energy range of the optical phonons at the ${\it \Gamma}$ point. These phonons have been identified for most HTSC,\cite{ThoCar} and the subtraction of the corresponding structures from the spectra has become a standard procedure to isolate electronic structures containing gap information. Electronic Raman scattering spectra are now available for many high-$T_c$ materials and, since they exhibit similar general features, most of these data are considered to be reliable. In this paper, we attempt to interpret these spectra from a theoretical point of view based on the full 3D one-electron band structure. We pay attention to both, line shapes and {\it absolute} scattering efficiencies. The theory of electronic Raman scattering in superconductors was pioneered by Abrikosov and coworkers in two important papers.\cite{AbrFal,AbrGen} In the first, they developed a theory for the scattering efficiency of {\it isotropic} Fermi liquids under the assumption that the attractive interaction between quasiparticles can be neglected. In the second paper, they extended this approach to anisotropic systems, introduced the effective mass vertex concept, and included Coulomb screening. The current form of the theory, developed mainly by Klein {\it et al.},\cite{KleDie} takes into account the attractive pairing interaction and emphasizes the role of gauge invariance as well as the polarization dependence for anisotropic gaps. In order to compare the theoretical predictions with the experiment, we evaluate them numerically in a quantitative manner (including {\it absolute} scattering efficiencies!) and compare them to the experimental findings. Several calculations of the electronic Raman scattering efficiency of HTSC have already been published. Some of them use highly simplified 2D band structures and a decomposition of the Raman vertex $\gamma_{\mbf k}$ in Fermi surface (FS) harmonics\cite{Allen} or Brillouin zone (BZ) harmonics, as well as FS integrations instead of the required BZ integrations.\cite{DevEin,DevRepl,DevEinPRB} The results of these calculations depend very strongly on the number of expansion coefficients used for $\gamma_{\mbf k}$ and their relative values. Another approach\cite{KraMaz} involves the use of band structures calculated in the framework of the local density approximation\cite{KohSha} (LDA) using the LMTO method.\cite{AndJep,AndLie} Within the approximations of the LDA, this Raman vertex is exact, i.e.\ the only errors made in such a calculation arise from limitations of the LDA method itself and from the discretization of the Brillouin zone or Fermi surface. Some of these calculations, however, suffer from the fact that only the imaginary part of the Tsuneto function\cite{Tsu} has been used, and that only 2D integrations were performed.\cite{KraCar} The present approach\cite{CarStr} is based on the full 3D LDA-LMTO band structure. It uses a BZ integration, screening effects are included, and both the real and imaginary part of the Tsuneto function are used as required by the theory. Electronic Raman spectra are calculated for ${\rm YBa}_2{\rm Cu}_3{\rm O}_7$ (Y-123) and ${\rm YBa}_2{\rm Cu}_4{\rm O}_8$ (Y-124). The orthorhombicity of the cuprates is also taken into account in the Raman vertex since we use as starting point the band structure of the {\it orthorhombic} materials. The cuprates under consideration are not only of interest because of their superconducting, but also of their strange normal-conducting properties. Usual metals should show peaks in their Raman spectra at their plasma frequencies corresponding to Raman shifts of a few ${\rm eV}} \def\meV{{\rm meV}$. The optimally doped cuprates, in contrast, show a very broad electronic background (from $0$ to about $1\,{\rm eV}} \def\meV{{\rm meV}$ Raman shift), which is almost independent of temperature and frequency. The spectra of the underdoped HTSC, such as Y-124, show some temperature dependence at low frequencies ($\hbar\omega\ll kT$). It is possible to explain these peculiarities, together with other properties, by assuming a certain form of the quasiparticle lifetime, as was done in the Marginal Fermi Liquid theory.\cite{Var,VarII} For the superconducting state, various forms for the gap function have been proposed. That which has received most experimental support has ${d_{x^2-y^2}}$ symmetry, i.e., $B_{1g}$ symmetry in tetragonal HTSC. The power of Raman scattering to confirm such gap function has been questioned, because, among other difficulties to be discussed below, it only probes the {\it absolute} value of the gap function, i.e.\ it cannot distinguish between a ${d_{x^2-y^2}}$-like gap function (for instance $\cos2\phi$), and a $\left|\cos2\phi\right|$ gap function, which corresponds to anisotropic $s$ ($A_{1g}$) symmetry. However, it was pointed out that addition of impurities can be used to effect the distinction.\cite{DevImp} This paper is organized as follows: in Sec.~\ref{sec:ldabands}, we review the main properties of the band structures of the investigated cuprates, as obtained by LDA-LMTO calculations. Sec.~\ref{sec:gtheory} discusses the theory of electronic Raman scattering in systems with an anisotropic band structure. We first introduce the basic concept of Raman vertex, and then present an expression relating the scattering efficiency to the Raman susceptibility. In Sec.~\ref{sec:stheory} and~\ref{sec:ntheory} we derive expressions for the Raman susceptibility in the superconducting and normal conducting phases, and discuss some effects not directly contained in the presented form of the theory. Section~\ref{sec:expts} is concerned with aspects of the experiments which have to be taken care of, especially with regard to the comparison with the theory. Finally, in Sec.~\ref{sec:numerics} the results of our numerical calculation are presented and compared to the experimental results. The difference between calculations involving only FS averaging, and those in which such averaging is performed over the whole BZ, are discussed. \section{LDA band structure} \label{sec:ldabands} The basis of our calculation is the LDA-LMTO band structure of the HTSC under consideration.\cite{AndLie} For the sake of further discussion, we shall describe briefly such band structure. The Fermi surface of $\rm YBa_2Cu_3O_7$ (Y-123)\cite{AndLie} consists of four sheets, an even and an odd $pd\sigma$-like plane band, a $pd\sigma$-like chain sheet and a very small $pd\pi$-like chain sheet. The latter is predicted by the full-potential LMTO calculations as well as LAPW calculations.\cite{OKAnd} We use the atomic spheres approximation (ASA) to the LMTO, which does not reproduce this rather small feature. In the case of Y-\nobreak123 the three $pd\sigma$-like conduction bands extend from $-1\,{\rm eV}} \def\meV{{\rm meV}$ to $2\,{\rm eV}} \def\meV{{\rm meV}$ relative to the Fermi energy. They are embedded in a broad valence band, which ranges from $-7\,{\rm eV}} \def\meV{{\rm meV}$ to $2\,{\rm eV}} \def\meV{{\rm meV}$ and consists of 36 bands (mainly Cu-$d$ and O-$p$ orbitals). Below $-7\,{\rm eV}} \def\meV{{\rm meV}$, there is a gap of $4\,{\rm eV}} \def\meV{{\rm meV}$. Above the conduction band, there is another gap of $0.5\,{\rm eV}} \def\meV{{\rm meV}$, above which are the lowest fully unoccupied bands which consist mainly of $d$ orbitals of Y and Ba. The band structure of $\rm YBa_2Cu_4O_8$ (Y-124) shows similar features. There is an additional $pd\sigma$-like chain band, while the $pd\pi$-like chain bands are predicted by both, full-potential LMTO and LAPW to contain no holes, i.e.\ to be completely filled. An interesting feature of the band structure of both Y-123 and Y-124 is an extended saddle point\cite{AndJep} on the $k_x$-axis near the $X$ point. This extended saddle point corresponds to a van Hove singularity at approximately $25\,\meV$ (Y-123) and $110\,\meV$ (Y-124), respectively, below the Fermi level. As will be shown, the comparatively large density of states in this energy region and the warped nature of the corresponding bands has an influence on the calculated electronic Raman spectrum. The band structure which we used in our numerical calculations was evaluated for Y-123 on a mesh of $48\times48\times12$ points in the first BZ, involving $4373$ irreducible points. The band structure of Y-124 is less sensitive to the resolution of the grid (because the extended saddle point lies deeper with respect to the Fermi surface). It was thus sufficient to use a $24\times24\times12$ mesh with $1099$ irreducible points. The calculations of the self-consistent potential have been performed in the ASA. Therefore, the $pd\pi$ chain band around the $S$ point, which should be partly filled, is completely filled. Because of the small number of states involved, we do not think this should affect significantly our results. As stressed above, our calculations are based on a band structure obtained within the LDA. We are aware of the fact, that the mean free path for transport in the direction of the $c$-axis is smaller than the size of the unit cell, i.e.\ that a description by means of a band structure $\epsilon_{n{\mbf k}}$ may be questionable (Ioffe-Regel-limit). Nevertheless a nontrivial band structure in the $c$-direction may simulate some of the $c$-direction confinement effects and represent, after integration along $k_z$, a reasonable 2D band structure. \section{General theory} \label{sec:gtheory} Two approaches have been used to derive the cross section (called scattering efficiency when referred to unit path length in the solid) of electronic Raman scattering in superconductors with anisotropic Fermi surfaces. The first uses Green's functions,\cite{AbrFal,AbrGen,KleDie,Kos} and the second the kinetic equation.\cite{Woe,DevEin} Both start with the simplification of the Hamiltonian, using ${\mbf k}\cdot{\mbf p}$ theory, which relates the Raman vertex $\gamma_{\mbf k}$ to the inverse effective mass tensor.\cite{AbrGen} We first briefly review this procedure and, subsequently, the derivation of the expression for the scattering efficiency using the diagrammatic approach. \subsection{The Raman vertex} To derive an expression for the efficiency for electronic Raman scattering, one has to replace the momentum ${\mbf p}$ in the Hamiltonian by ${\mbf p}-(e/c){\mbf A}$. This yields two distinct perturbation terms: $H_{AA}=(r_0/2){\mbf A}^2$, quadratic in the vector potential, and $H_A=-(e/mc){\mbf A}\cdot{\mbf p}$, linear in $A$ (we use the transverse gauge; $r_0=e^2/(m_0c^2)$ denotes the classical electron radius). The relevant states in the theory are composed of the state of electrons in the sample plus the state of the photon field. The initial state of the photon field has $n_L$ laser photons with wave vector ${\mbf k}_L$ and polarization ${\mbf e}_L$ and $n_S=0$ scattered photons with wave vector ${\mbf k}_S$ and polarization ${\mbf e}_S$. The final state has one laser photon less but one scattered photon (Stokes scattering). The vector field can thus be written as a superposition of an incoming and a scattered plane wave, ${\mbf A}={\mbf A}_L+{\mbf A}_S$ with \begin{equation} {\mbf A}_S=A_S^+{\mbf e}_S^* e^{-i{\mbf k}_Sr}~,\qquad {\mbf A}_L=A_L^-{\mbf e}_L e^{i{\mbf k}_Lr}~, \end{equation} where $A_S^+$ contains the creation operator for the scattered photon and $A_L^-$ the annihilation operator for the laser photons (note that these are not hermitian). Since Raman scattering is a second order process in ${\mbf A}$, the term $H_{AA}$ has to be treated in first order perturbation theory. It is therefore nonresonant and includes only intraband scattering. The matrix elements are given by \begin{eqnarray} M^{(1)}_{n_fn_i}({\mbf q},{\mbf k}) &=& {1\over2}r_0\braket{n_f{\mbf k}+{\mbf q}}{{\mbf A}^2}{n_i{\mbf k}} \nonumber\\ &=& r_0 \expv{A_S^+A_L}{\mbf e}_S^*{\mbf e}_L\delta_{n_in_f}~, \label{fstpb} \end{eqnarray} whereas $\expv{A_S^+A_L}$ denotes a matrix element involving the initial and final state of the photon field and ${\mbf q}={\mbf k}_L-{\mbf k}_S$ is the momentum transfer from the photon field to the sample. For other values of ${\mbf q}$, the matrix element $(\ref{fstpb})$ vanishes. The second term, $H_A$, produces resonances via second order perturbation theory. It has the form \begin{equation} M^{(2)}_{n_fn_i}({\mbf q},{\mbf k}) = r_0 \expv{A_S^+A_L} \sum_{n_m} \Gamma^{(2)}_{n_fn_i;n_m}({\mbf q},{\mbf k}) \label{sndpb} \end{equation} with the expression \begin{eqnarray} \Gamma^{(2)}_{n_fn_i;n_m}({\mbf q},{\mbf k}) & = & \label{second}\\ \noalign{\hskip0.2\baselineskip}% &&\hspace{-2.5cm} {\braket{n_f{\mbf k}+{\mbf q}}{{\mbf e}_S^*{\mbf p}}{n_m{\mbf k}+{\mbf k}_L} \braket{n_m{\mbf k}+{\mbf k}_L}{{\mbf e}_L{\mbf p}}{n_i{\mbf k}} \over\epsilon_{n_i{\mbf k}}-\epsilon_{n_m{\mbf k}+{\mbf k}_L}+\omega_L+i0} + \nonumber \\ &&\hspace{-2.5cm} + {\braket{n_f{\mbf k}+{\mbf q}}{{\mbf e}_L{\mbf p}}{n_m{\mbf k}-{\mbf k}_S} \braket{n_m{\mbf k}-{\mbf k}_S}{{\mbf e}_S^*{\mbf p}}{n_i{\mbf k}} \over\epsilon_{n_i{\mbf k}}-\epsilon_{n_m{\mbf k}-{\mbf k}_S}-\omega_S+i0}~. \nonumber \end{eqnarray} Here, $\omega_L$ and $\omega_S$ are the frequency of the incoming and scattered light, respectively. Note that the states in the sum above are states of the sample only. We have used Bloch states with band and crystal momentum indices. The wavevectors of light, ${\mbf k}_L$ and ${\mbf k}_S$, can usually be neglected in the matrix elements of Eq.~$(\ref{second})$ because $v_F\ll c$. For the same reason, $\epsilon_{n_m,{\mbf k}+{\mbf k}_L}\approx\epsilon_{n_m,{\mbf k}}$. Therefore, we introduce the symbol $\Gamma^{(2)}_{n_fn_i;n_m}({\mbf k})$ to denote expression~$(\ref{second})$ with the light wavevectors set equal to zero. If we now add the contributions of both terms in Eqs.~(\ref{fstpb}) and~(\ref{sndpb}) and introduce second quantization, we are left with the {\it effective Hamiltonian} \begin{equation} H_R = r_0\expv{A_S^+A_L}\,\tilde\rho_{\mbf q} \label{effham} \end{equation} as perturbation Hamiltonian leading to Raman scattering. The effective density operator $\tilde\rho_{\mbf q}$ can be expressed in the form \begin{equation} \tilde\rho_{\mbf q} = \sum_{n_f,n_i,{\mbf k}} \gamma_{n_fn_i}({\mbf k})\, c^+_{n_f,{\mbf k}+{\mbf q}} c_{n_i,{\mbf k}}~. \label{flucop} \end{equation} using fermionic creation and annihilation operators for Bloch electrons as well as the nondiagonal Raman vertex \begin{equation} \gamma_{n_fn_i}({\mbf k}) = {\mbf e}_S^*{\mbf e}_L\delta_{n_fn_i} + \sum_{n_m} \Gamma^{(2)}_{n_fn_i;n_m}({\mbf k})~. \label{raver} \end{equation} If we are interested mainly in the low frequency region, say Raman shifts below $50\,\meV$, no {\it real} interband transitions of significant weight are possible. This can easily be seen from the band structure (Fig.~2 of Ref.~\onlinecite{AndJep}). Therefore, we introduce the (intraband) {\it Raman vertex} $\gamma_n({\mbf k})=\gamma_{nn}({\mbf k})$. We proceed by discussing a very important simplification of $(\ref{second})$ (with $n_i=n_f=n$ and ${\mbf q}=0$), the {\it effective mass approximation}. Four different cases will be discussed. First, the virtual intraband transition with $n_m=n_i$. In this case, up to first order in $v_F/c$, we have $\braket{n_m{\mbf k}}{{\mbf p}}{n_i{\mbf k}} = \braket{n_i{\mbf k}}{{\mbf p}}{n_m{\mbf k}}$ (remember that $n_i=n_f$) and $\epsilon_{n_i{\mbf k}}-\epsilon_{n_m{\mbf k}}=0$. Then, it can be seen that the contributions of virtual {\it intra\/}band transitions relative to the contribution of virtual {\it inter\/}band transitions to intermediate states are of the order of the Raman shift over the laser frequency, i.e.\ $\omega/\omega_L\ll1$, and can therefore be neglected. The second case are the virtual {\it inter\/}band transitions involving bands which are much farther away from the FS than the light frequency. Then, because of $\abs{\epsilon_{n_i}-\epsilon_{n_m}}\gg\omega_L$, the light frequencies $\omega_L$ as well as $\omega_S$ can be neglected in $(\ref{second})$. The third case also involves virtual interband transitions, but for bands at about the laser frequency above the Fermi surface. Here, the scattering is resonant, and the spectra are expected to depend strongly on the laser wavelength. One can try to avoid this situation by using different laser lines. So we assume that in the third case $\omega_L$ and $\omega_S$ also can be neglected. Finally, the forth case consists of virtual interband transitions to neighboring bands with $\Delta\epsilon\ll\omega_L$. In this case, neglecting $\omega_L$ and $\omega_S$ is more difficult to justify. We do it nevertheless and reach the approximate conclusion that we can neglect the light frequencies in Eq.~(\ref{second}) and can restrict the sum in $(\ref{sndpb})$ to all $n_m\not=n_i$. Then, Eq.~(\ref{raver}) becomes completely equivalent to the expression for the inverse effective mass from ${\mbf k}\cdot{\mbf p}$ theory and we can write \begin{equation} \gamma_n({\mbf k})={m\over\hbar^2} \sum_{i,j} {\mbf e}^*_{S,i} {\partial^2\epsilon_{n{\mbf k}}\over\partial k_i\partial k_j} {\mbf e}_{L,j} \end{equation} i.e.\ the Raman vertex is equal to the inverse effective mass contracted with the polarization vectors of the laser light and the scattered light, respectively. Therefore, using the term $H_R$ with the intraband Raman vertex $\gamma_{n{\mbf k}}=\gamma_{nn}({\mbf k})$ in Eq.~(\ref{flucop}) as perturbation to the Hamiltonian for ${\mbf A}=0$ and treating this in first order perturbation theory is, under the mentioned restrictions, equivalent to taking into consideration both terms $H_A$ and $H_{AA}$.\cite{AbrGen} According to the LMTO calculations, for Y-123 and Y-124 there are bands above a band gap between approximately $2\,{\rm eV}} \def\meV{{\rm meV}$ and $2.5\,{\rm eV}} \def\meV{{\rm meV}$ above the Fermi energy. These bands can present a problem with respect to the former discussion, because they are almost resonant for typical laser wavelengths like $514.5\,{\rm nm}$. The same is true for the conduction bands, which extend until $2\,{\rm eV}} \def\meV{{\rm meV}$ above the Fermi surface. Note that due to the strong on-site interaction at the Cu-$d$ orbitals, correlation effects are expected to be important in the electronic structure. It is possible that at energies of the order of $1\,{\rm eV}$ or more above the Fermi surface the picture of the Hubbard bands is a better description of the band structure and may explain the weak dependence of the Raman spectra on the laser frequency observed for laser frequencies in the visible range. The band structure shows many bands at about the laser frequency below the Fermi energy. These should yield resonant contributions to the Raman efficiency. Because the Raman vertex $\gamma_{\mbf k}$ is, in the given approximation, the second derivative of the energy with respect to ${\mbf k}$, the $A_{2g}$ component for tetragonal crystals vanishes in this version of the theory ($A_{2g}$ is the symmetry of an antisymmetric tensor). If one considers once more the effects of a nearby resonance, it can be easily seen that the Raman tensor does not have to be symmetric. This stresses again the questionability of the effective mass approach if the scattering is resonant. \subsection{The scattering efficiency} Using the effective mass approach, we arrived at the effective Hamiltonian $(\ref{effham})$ with the effective mass determining the Raman vertex. This effective Hamiltonian is linear in ${\mbf A}_L\cdot{\mbf A}_S$. The derivation of the scattering efficiency using linear response theory is now a straightforward task. The first step is finding a relation between the Raman efficiency and a dynamical structure factor of the sample. Then, in a next step, the fluctuation-dissipation theorem is used to connect the dynamical structure factor to the imaginary part of a susceptibility, in our case the {\it Raman susceptibility}. To establish the relation to the dynamical structure factor, we add the time evolution factor $e^{-i\omega t}$ to the effective Hamiltonian $(\ref{effham})$ and use the golden rule to find the transition rate from a state $i$ to a state $f$ of the sample. Then, we sum over all final states $f$ of the sample and do a thermal averaging over the initial states $i$. The transition rate from a state with $n_L\equiv n_{{\mbf k}_L{\mbf e}_L}$ laser photons and no scattered photon to a state with $n_L-1$ laser photons and $n_S\equiv n_{{\mbf k}_S{\mbf e}_S}=1$ scattered photon at a temperature $T$ is given by the expression \begin{equation} \Gamma^T({\mbf k}_L,{\mbf e}_L;{\mbf k}_S,{\mbf e}_S) = {2\pi\over\hbar} r_0^2\cdot \absii{\expv{A_S^+A_L}}\cdot \tilde S^T({\mbf q},\omega) \end{equation} (the superscript $T$ denotes temperature dependency) whereas \begin{eqnarray} \tilde S^T({\mbf q},\omega) &=& \nonumber\\ &&\hspace{-1cm}\sum_{i,f}{e^{-\beta E_i}\over\cal Z} \absii{\braket{f}{\tilde\rho_{\mbf q}}{i}} \delta(E_f-E_i+\hbar\omega) \end{eqnarray} is a {\it generalized dynamical structure factor} (of the sample!). The partition function is denoted by ${\cal Z}$, and $\beta$ is the inverse temperature. Now, we sum over all final states in a certain region $d\Omega\,d\omega_S$ of $k$-space around ${\mbf k}_S$ and normalize to the incoming flux $\hbar cn_L$. This yields the expression \begin{equation} {d^2\sigma\over d\Omega\,d\omega}({\mbf q},\omega) = {\omega_S\over\omega_L} r_0^2 \tilde S^T({\mbf q},\omega) \label{rameff} \end{equation} for the differential cross section $d^2\sigma/d\Omega d\omega$ for a given Raman shift $\omega$ and a given momentum transfer ${\mbf q}$. This differential cross section is proportional to the scattering volume. When performing the calculation for a scattering volume equal to unity, $\sigma$ becomes the commonly used {\it Raman scattering efficiency}. Finally, one can define a linear response function, the {\it Raman susceptibility} \begin{equation} \chi_{\rm Raman}({\mbf q},t) = {i\over\hbar} \mathop{\rm Tr}\{{\cal Z}^{-1}e^{-\beta H_0} [\tilde\rho_{\mbf q}(t),\tilde\rho_{-{\mbf q}}(0)]\} \label{RamSus} \end{equation} and its Fourier-transformed $\chi_{\rm Raman}({\mbf q},\omega)$. To relate the imaginary part of this quantity to the structure function $\tilde S^T({\mbf q},\omega)$, we use the fluctuation-dissipation theorem. The result is \begin{equation} \tilde S^T({\mbf q},\omega) = -{1\over\pi}(1+n_\omega) \mathop{\rm Im}\nolimits\chi_{\rm Raman}({\mbf q},\omega) \label{fdt} \end{equation} with the Bose factor $n_\omega$. Equations $(\ref{rameff})$ and $(\ref{fdt})$ relate the Raman efficiency directly to the imaginary part of the Raman susceptibility. This evaluation of the Raman susceptibility shall be given separately, {\rm (i)} in Sec.~\ref{sec:stheory} for the superconducting phase and Raman shifts of the order of the gap, and {\rm (ii)} in Sec.~\ref{sec:ntheory} for large Raman shifts in the superconducting phase and for the normal phase. \section{Theory: Superconducting phase} \label{sec:stheory} As pointed out in Ref.~\onlinecite{KleDie}, the Raman susceptibility due to pair-breaking and including screening is given by a polarization-like bubble made of a renormalized Raman vertex $\Lambda_{\mbf k}$, a Raman vertex $\gamma_{\mbf k}$, and in between two Green's function lines for Bogoliubov quasiparticles (Fig.~\ref{screening}a). The vertex renormalization includes corrections for Cooper-pair-producing attractive interaction as well as the repulsive Coulomb interaction, the Dyson equation for the vertex $\Lambda_{\mbf k}$ in the limit ${\mbf q}\to0$ is given by Fig.~13 in Ref.~\onlinecite{KleDie}. To show more clearly the effect of screening, we write the equation for the Raman susceptibility as given in Fig.~\ref{screening}b and~\ref{screening}c. Figure~\ref{screening}b (with $a=\gamma_{\mbf k}$ and $b=\gamma_{\mbf k}$) shows the unscreened susceptibility $\chi_{\gamma\gamma}$ given by a bare polarization bubble with two Raman vertices $\gamma_{\mbf k}$ and the contraction of a BCS-like ladder sum with two Raman vertices. Therefore, $\chi_{\gamma\gamma}$ includes the attractive Cooper-pair-producing interaction. We include Coulomb screening by virtue of a RPA-like sum given in Fig.~\ref{screening}c. The effect of screening on the electronic Raman scattering can now easily be seen.\cite{AbrGen} If we denote by $\chi_{ab}$ a bubble, renormalized by pairing interaction, with vertices $a$ and $b$ at the ends as in Fig.~\ref{screening}b, the RPA-chain can be easily summed up (see Fig.~\ref{screening}c) yielding \begin{equation} \chi_{\rm Raman}({\mbf q}\to0,\omega) = \chi_{\gamma\gamma}(\omega) - {\chi_{\gamma 1}^2(\omega)\over\chi_{11}(\omega)}~, \label{chir} \end{equation} where terms of order ${\mbf q}^2$ have been dropped. In Eq.~$(\ref{chir})$ we have used the fact that $V_{\mbf q}/(1-\chi_{11}V_{\mbf q})$ equals $-\chi_{11}^{-1}(1-1/\varepsilon)$, and the factor $(1-1/\varepsilon)$ is $1+O({\mbf q}/q_{TF})^2$. Without taking into account Coulomb interaction, the Green's functions have a well-known massless pole (Goldstone mode) which is a consequence of the breaking of gauge symmetry in the superconducting phase.\cite{And} Coulomb interaction makes this pole acquire a finite mass (which can be shown to correspond to the plasma frequency), so if we correctly include Coulomb screening (not done in Ref.~\onlinecite{BraCar}) we no longer have a Goldstone mode, but a massive Anderson-Bogoliubov mode. This mode has the energy $\hbar\omega_p$ ($\omega_p$ is the plasma frequency) at the ${\it\Gamma}$ point and is therefore negligible for the low energy behavior of the Raman spectra. The susceptibilities $\chi_{ab}$ in Fig.~\ref{screening}b are like a ladder sum contracted with vertices $a_{\mbf k}$ and $b_{\mbf k}$ and can be written as a sum \begin{equation} \chi_{ab}({\mbf q}{=}0, \omega) = \sum_{\mbf k} a_{\mbf k} b_{\mbf k} \lambda_{\mbf k}(\omega) \end{equation} which involves the Tsuneto function\cite{Tsu} $\lambda_{\mbf k}(\omega)$. For small values of ${\mbf q}$ (compared to the inverse coherence length $\xi$ and the Fermi wave vector $k_F$), the attractive interaction does not have to be taken into account in the summation of the ladder, and the Tsuneto function is given simply by a unmodified bubble and can be evaluated easily to be \begin{eqnarray} \lambda_{\mbf k}(\omega) &=& {\Delta^2_{\mbf k}\over E^2_{\mbf k}} \tanh\left(E_{\mbf k}\over 2T\right)\times\nonumber\\ &&\times\left({1\over2E_{\mbf k}+\omega+i0}+{1\over2E_{\mbf k}-\omega-i0}\right)~. \label{tsufct} \end{eqnarray} Equation~$(\ref{tsufct})$ involves the gap function $\Delta_{\mbf k}$ (which depends on the temperature) and the quasiparticle dispersion relation $E_{\mbf k}^2=\xi_{\mbf k}^2+\Delta_{\mbf k}^2$ with $\xi_{\mbf k}^2=(\epsilon_{\mbf k}-\epsilon_F)^2$. The constants $\hbar$ and $k_B$ have been set equal to $1$. As already mentioned, vertex corrections due to the pairing interaction are neglected. This approximation is valid for $q\ll\xi^{-1},k_F$ (Ref.~\onlinecite{DevEinPRB}) and $\omega\ll\omega_p$, because the Anderson-Bogoliubov pole at the plasma frequency need no longer be included. A first and very important fact in the expressions above is that they contain only the absolute square of the gap function, i.e.\ Raman scattering is {\it not phase sensitive}, and consequently cannot distinguish between a strongly anisotropic $s$ gap $\left|{d_{x^2-y^2}}\right|$ and a ${d_{x^2-y^2}}$ gap. In the preceding calculation of the unscreened correlation functions $\chi_{ab}$, we have neglected impurity scattering as well as scattering between quasiparticles (collisionless regime). In isotropic $s$-wave superconductors at $T=0$ and for Raman shifts $\omega\ll2\Delta$, it is perfectly reasonable to neglect impurity scattering, because in this regime pair breaking is not possible.\cite{AndImp} Also, the scattering between quasiparticles can be neglected because their density is very small for small temperatures $T\ll T_c$. For $d$-wave superconductors this is no longer true. The effect of impurities will be discussed in the next subsection, whereas a discussion about scattering between quasiparticles can be found in Sec.~\ref{sec:ntheory}. The second term of $(\ref{chir})$, representing screening, vanishes if the average of $\gamma_{\mbf k}\cdot\lambda_{\mbf k}$ does. The Tsuneto function is fully symmetric, i.e.\ has $A_{1g}$ ($D_{4h}$ group) or $A_g$ ($D_{2h}$) symmetry regardless of gap symmetry. As a consequence, the screening term vanishes unless the Raman vertex has the same symmetry as the crystal. In the tetragonal case, $A_{1g}$-like vertices are screened, but $B_{1g}$- and $B_{2g}$-like are not. This is different for orthorhombic HTSC of the YBCO-type. In this case the Tsuneto function has $A_1$ symmetry, and the same is true for the ${d_{x^2-y^2}}$-like component of the mass ($B_{1g}$ of $D_{4h}$ group, $A_g$ of $D_{2h}$). Consequently, in these orthorhombic crystals, the $B_{1g}$ component is also screened. This discussion is also applicable to BISCO, but with interchanged roles of $B_{1g}$ and $B_{2g}$ modes because of the different orientation of the crystallographic unit cell with respect to the Cu-O bonds. In tetragonal systems, the $B_{1g}$ component of the Raman vertex has nodes at the same position as the gap function. This has severe consequences for the low-energy part of the spectra.\cite{DevEinPRB} In two dimensions, the existence of the nodes of the gap function in the case of a ${d_{x^2-y^2}}$ gap results in a linear density of states at low energies. If the vertex has a finite value in this region, the imaginary part of the Raman susceptibility is also linear in the frequency. If the vertex has a node, however, its magnitude squared becomes quadratic with respect to the gap on the Fermi surface. This causes two additional powers of the frequency to appear, the $B_{1g}$ component of the scattering efficiency is cubic at low frequencies.\cite{DevEin} Two effects can alter this behavior: an orthorhombic distortion and impurities. In our calculations, we focus on a ${d_{x^2-y^2}}$-like gap function which is only a function of the direction in $k$-space, but not of the magnitude of ${\mbf k}$, since the values of the gap functions sufficiently far from the Fermi surface do not affect the results. We are using the same gap function for all bands involved. \subsection{Effect of impurities} \label{sec:stheory:imp} In contrast to scattering at non-magnetic impurities in conventional (isotropic) superconductors, the influence of impurity scattering plays an important role for superconductors with anisotropic gaps and its effect on the Raman spectrum is most pronounced for superconductors which exhibit regions in $k$-space where the gap almost or completely vanishes. It was shown\cite{BorHir,DevImp} that in the case of $d$-wave pairing, impurity scattering can be described by extending the nodal points on the 2D FS to small finite regions with vanishing gap. This causes a nonvanishing density of states at the Fermi energy. For anisotropic $s$-wave pairing the gap anisotropy becomes smeared out leading to an increase of the minimum gap value $\Delta_{\rm min}$. In the case of a $\abs{{d_{x^2-y^2}}}$ gap, this minimum gap increases monotonically with the impurity concentration $n_i$ for small values of $n_i$. The renormalization of the gap function by the presence of impurities causes an additional contribution, which is linear in the Raman shift $\omega$ for small Raman shifts $\omega$, in the Raman spectra.\cite{DevImp} This has consequences for the $B_{1g}$ spectrum of a {\it tetragonal} crystal, which, according to the theory, has a cubic $\omega$-dependence, because a linear frequency dependence is added. As will be discussed in the next subsection, the orthorhombicity of the YBCO compounds also causes a linear addition to the cubic behavior of the $B_{1g}$ channel spectrum. In the case of a $\abs{{d_{x^2-y^2}}}$-like, $A_g$ symmetry gap function the impurity-induced minimal gap $\Delta_{\rm min}$ causes an excitation-free region to show up in the electronic Raman spectrum below a Raman shift of $2\Delta_{\rm min}$. \subsection{Effect of orthorhombic distortion} \label{sec:stheory:odist} As already mentioned, orthorhombic distortions, i.e.\ deviations from the tetragonal symmetry, have a different effect on Y-123 and on Bi-2212. Consider the $B_{1g}\,(D_{4h})$ component of the inverse mass tensor in a {\it tetragonal} high-$T_c$ superconductor with a ${d_{x^2-y^2}}$-like gap. The $B_{1g}\,(D_{4h})$ mass has its nodes in directions diagonal to the axes of the copper planes; the same is true for the gap function. As mentioned above, this results in the $\omega^3$-dependence of the Raman efficiency for $B_{1g}\,(D_{4h})$ scattering, in contrast to the $\omega$-dependence predicted for $A_{1g}$ and $B_{2g}$ scattering. Let us now consider the orthorhombic distortion present in Y-123. The zeros of the $B_{1g}\,(D_{4h})$ mass shift because there are no longer mirror planes through the $(110)$ axes. For this reason, the low-energy part of the spectrum acquires a linear component in addition to the $\omega^3$ component of the $D_{4h}$ case. In Bi-2212 the situation is different because the orthorhombic crystallographic cell is rotated by $45^\circ$ with respect to the $a$- and $b$-axes: the orthorhombic distortion preserves the mirror planes $[a\pm b,c]$. Consequently, the $B_{1g}$ zeros stay at the same position, the low-energy efficiency acquires no linear component. \subsection{Effect of multilayers} \label{sec:stheory:mlayer} In systems with one layer of Cu-$\rm O_2$ planes per unit cell there is only one sheet of Fermi surface and the mass fluctuations are essentially intraband mass fluctuations, which are very sensitive to the scattering polarizations. The scattering related to the average mass is fully screened. The simplest $A_{1g}\,(D_{4h})$ scattering is related to a Raman vertex of the form $\cos4\phi$ symmetry while $B_{1g}\,(D_{4h})$ scattering is obtained for a $\cos2\phi$ vertex. In multilayer systems, interband fluctuations between the various sheets FS are also important. The lowest component of such fluctuations corresponds to different {\it average} masses in each FS sheet. Such fluctuations do not depend on the scattering polarizations and lead to unscreened scattering of $A_g$ symmetry. \subsection{Effect of sign change of $\gamma_{\mbf k}$ on the Fermi surface} \label{sec:stheory:signch} The behavior of the Raman vertex near the Fermi surface, especially its sign, is crucial for the scattering efficiency and, in particular, for the effect of screening. {\it Antiscreening}, i.e.\ an {\it enhancement} of the scattering efficiency by screening, can occur if the Raman vertex changes sign on the Fermi surface. This can be seen by considering the screening part \begin{equation} \mathop{\rm Im}\nolimits{\chi_{\rm Scr}} = - \mathop{\rm Im}\nolimits{\chi_{\gamma1}^2\over\chi_{11}} \label{scrpart} \end{equation} of the Raman susceptibility. A positive value of $\mathop{\rm Im}\nolimits{\chi_{\rm Scr}}$ enhances the efficiency, i.e.\ corresponds to antiscreening. To show how antiscreening arises, we first write the screening term $\mathop{\rm Im}\nolimits\chi_{\rm scr}$ in terms of the real and imaginary parts ${\lambda'}\equiv\mathop{\rm Re}\nolimits\lambda$ and ${\lambda''}\equiv\mathop{\rm Im}\nolimits\lambda$ of the Tsuneto function and the Raman vertex $\gamma$ as \begin{equation} \mathop{\rm Im}\nolimits{\chi_{\rm Scr}}={ {\expv{\gamma{\lambda'}}}^2\expv{{\lambda''}}-{\expv{\gamma{\lambda''}}}^2\expv{{\lambda''}} -2\expv{\gamma{\lambda'}}\expv{\gamma{\lambda''}}\expv{{\lambda'}}\over {\expv{{\lambda'}}}^2+{\expv{{\lambda''}}}^2}~.\label{screefo} \end{equation} The imaginary part of the Tsuneto function ${\lambda''}$ is a positive $\delta$-function. Consequently, the quantity $\expv{{\lambda''}}$ is a positive function of the Raman shift $\omega$. If $\gamma_{\mbf k}$ changes sign in a region around the Fermi surface, it is possible that $\expv{\gamma{\lambda''}}$ changes sign as a function of $\omega$, i.e.\ has a zero. At the position of this zero, the second and the third term in the numerator of $(\ref{screefo})$ vanish. The first term, ${\expv{\gamma{\lambda'}}}^2\expv{{\lambda''}}$, is positive and can become dominant in Eq.~(\ref{screefo}). In this case antiscreening results. In the Appendix~A will be shown that antiscreening is particularly sensitive to the sign of the Raman vertex on parts of the Fermi surface around the directions of the nodes of the gap function $\Delta_{\mbf k}$. \section{Theory: Normal phase} \label{sec:ntheory} In the normal phase, the exact mechanism which produces a finite Raman intensity almost constant over a broad frequency and temperature range, is not known. Therefore, we assume some scattering mechanism, which implies a finite lifetime of the quasiparticles. Candidates for this scattering are the quasiparticle-quasiparticle scattering in Marginal Fermi Liquid theory\cite{Var} (MFL), impurity scattering\cite{HirWoe} or scattering due to spin fluctuations.\cite{QuiHir} A self energy with non-vanishing imaginary part yields a susceptibility of the form \begin{equation} \chi_{ab}({\mbf q}{=}0, \omega)=\sum_k a_{\mbf k} b_{\mbf k}\nu_{\mbf k}(\omega) \end{equation} with the relaxation kernel (the function $f'$ is the derivative of the Fermi function with respect to the energy) \begin{equation} \nu_{\mbf k}(\omega) = -f'(\xi_{\mbf k}) {i\Gamma_{\mbf k}\over\omega+i\Gamma_{\mbf k}} \end{equation} and its imaginary part \begin{equation} \mathop{\rm Im}\nolimits\nu_{\mbf k}(\omega) = -f'(\xi_{\mbf k}) {\omega\Gamma_{\mbf k}\over\omega^2+\Gamma_{\mbf k}^2}~. \label{imnu} \end{equation} This can easily be seen by evaluating a bubble with two Greens function lines for quasiparticles with an imaginary part $\Gamma_{\mbf k}$ of the self energy. Note that in the superconducting phase for Raman shifts larger than $\sim\!\Delta$, the relaxation effects described by (\ref{imnu}) are also of importance. The relevant relaxation kernel in this case is \begin{equation} \nu_{\mbf k}(\omega) = -f'(E_{\mbf k}) {\xi_{\mbf k}^2\over E_{\mbf k}^2} {i\Gamma_{\mbf k}\over\omega+i\Gamma_{\mbf k}}~, \label{drude} \end{equation} where $\xi_{\mbf k}^2=(\epsilon_{\mbf k}-\epsilon_F)^2$. To describe the constant background in the Raman spectra in the normal phase, one has to adopt the quasiparticle scattering rate of the MFL theory\cite{Var,VarII} \begin{equation} \Gamma_{\mbf k}(\omega) \sim \max(\alpha T, \beta\omega)~. \label{mfl} \end{equation} In order to evaluate the real part of $\nu_{\mbf k}$ using causality arguments, and to prevent divergences, we introduce a high-frequency cutoff $\omega_C$. Note that the nearly antiferromagnetic Fermi liquid\cite{MMP,BarPin} (NAFL) and also the nested Fermi liquid\cite{RuvVir} (NFL) yield a very similar quasiparticle scattering rate. The former can also provide a mechanism, which accounts for ${d_{x^2-y^2}}$ pairing. Similar results are obtained with Luttinger liquid based results.\cite{CarStr} Equation~(\ref{mfl}) yields a scattering continuum which is constant for frequencies smaller than $\min(\alpha T/\beta,T)$ and for frequencies larger than the temperature $T$, but with different intensities. In the first case, $\Gamma_{\mbf k}$ is proportional to the temperature, i.e.\ $\mathop{\rm Im}\nolimits\chi\sim\omega/T$. Multiplying by the Bose factor $1+n_\omega\sim T/\omega$ a constant is found. In the second case, $\Gamma_{\mbf k}\sim\omega$, and, consequently, $\mathop{\rm Im}\nolimits\chi={\rm const}$. The Bose factor is also constant and one is left with a constant Raman intensity. Note that in the first case, $\mathop{\rm Im}\nolimits\chi$ cancels the $\omega$- and $T$-dependence of the Bose factor. It has been shown,\cite{DonKir,TZhou,VarII} that ${\rm Y}{\rm Ba}_2{\rm Cu}_4{\rm O}_8$ does not exhibit this behavior. This has been attributed to the breakdown of MFL theory for not optimally doped cuprates.\cite{VarII} Actually, in this case the spectra are nearly temperature independent {\it after} dividing them by the Bose factor. We shall address this question once more at the end of this section. To discuss quasiparticle-quasiparticle (qp-qp) scattering, and its influence on electronic Raman scattering, we start with the case of a ${d_{x^2-y^2}}$ gap. Suppose the nodes of this gap have a width $\delta_0$ in $k$-space on the Fermi surface due to impurity scattering. We use the model of Eq.~(\ref{drude}) with a quasiparticle scattering rate $\Gamma_{\mbf k}$ independent of ${\mbf k}$ and discuss first the case $T=0$. Then it can be seen that the contribution of qp-qp scattering to the imaginary part of the Raman susceptibility $(\ref{RamSus})$ for low frequencies $\omega\ll\Delta_{\rm max}$ is proportional to the Drude-like factor $\omega\Gamma/(\omega^2+\Gamma^2)$ (which is, for small $\omega$ and low temperatures $T<\omega$, linear in $\omega$ if $\Gamma=\hbox{const}$ (semiconductors) or $\Gamma\sim\max(\omega^2,T^2)$ (FL), but constant as a function of $\omega$ if $\Gamma\sim\max(\omega,T)$ (MFL). In the tetragonal case, it is also proportional to the density of states at the Fermi surface and in the case of $A_{1g}$ and $B_{2g}$ polarizations to the width $\delta_0$, and in the case of $B_{1g}$ to the third power $\delta_0^3$ of the width $\delta_0$. The discussion for BISCO is analogous with the exception that $B_{1g}$ and $B_{2g}$ exchange their role. Finite, but small temperatures $T\ll\Delta_{\rm max}$ have the effect of enlarging the widths $\delta_0$ linearly in temperature, i.e.\ the temperature dependence of the contribution from qp-qp scattering is proportional to ${\rm const}+T$. Note that for $T\appgeq0$, the Bose factor changes the linear-in-$\omega$ dependence to a constant. For the anisotropic $s$ gap of the form $\abs{{d_{x^2-y^2}}}$ which acquires a finite minimum gap $\Delta_{\rm min}$ due to the presence of impurities,\cite{DevImp} the situation is different. The frequency dependence is also given by the factor $\omega\Gamma/(\omega^2+\Gamma^2)$ in addition to the Bose factor. But the temperature dependence is different. For temperatures $T\ll\Delta_{\rm min}$ smaller than the minimal gap, the density of quasiparticles is proportional to $\exp(-\Delta_{\rm min}/kT)$, i.e.\ the contribution of qp-qp scattering to the Raman efficiency is exponentially small. At $kT\approx\Delta_{\rm min}$, this exponential dependence on $T$ crosses over to a power law. The background electronic Raman spectrum in the normal phase is almost independent of temperature for nearly optimally doped high-$T_c$ compounds only. In the overdoped and underdoped case, the materials seem to show Fermi liquid-like behavior concerning the quasiparticle scattering rate $\Gamma_{\mbf k}$ (for small $\omega$).\cite{VarII,TZhou} The temperature dependence of the scattering rate $\Gamma_{\mbf k}$, as defined in $(\ref{drude})$, has been measured\cite{HacNem} for optimally doped and overdoped Bi-2212, and, especially in the case of the $B_{2g}\,(D_{4h})$ mode, the optimally doped sample shows $\Gamma=\alpha T$, whereas for the overdoped sample $\Gamma=\alpha'T^2+\Gamma_0$. Therefore, the overdoped sample shows properties of a normal Fermi liquid which are predicted by theory to have $\Gamma\sim\max(\omega^2,T^2)$. The $B_{1g}\,(D_{4h})$ mode result for the optimally doped sample yields the puzzling quasiparticle scattering rate $\Gamma={\rm const}$. \section{Experimental spectra} \label{sec:expts} The experimental determination of {\it absolute} Raman scattering intensities is plagued by a number of difficulties (a reason why usually ``relative units'' are found in the literature). The first is related to the presence of elastically scattered light in the spectra, in particular when non-ideal sample surfaces are involved. Depending on the quality of the spectrometer this leads to contributions extending typically, for the parameters of the present work, up to $50\,{\rm cm}^{-1}$ from the center of the laser line. These contributions can be filtered out using a premonochromator or notch filters but, in any case, Raman scattering measurements below $50\,{\rm cm}^{-1}$ remain difficult. The measurements discussed here have been performed by comparison with the known efficiency of silicon after correcting for differences in the scattering volumes. The procedure leads to errors of about 50\%. We use for comparison with the calculation the experimental data of Krantz {\it et al.}\cite{KraCar} in the case of Y-123, and Donovan {\it et al.}\cite{DonKir} in the case of Y-124. Our Figs.~\ref{Y123exp} and~\ref{Y124exp} are taken from these publications. In the case of Fig.~\ref{Y123exp} we have corrected a scale error in the abscissa found in Ref.~\onlinecite{KraCar}. In the case of Fig.~\ref{Y124exp} we have calculated the $A_{1g}$ component from the experimental results for the $(x'x')$ and $(xy)$ polarizations. The classification of the measured spectra according to irreducible representations of the symmetry group of the crystal is performed with the use of the Raman tensor $\hat R$ which is related to the Raman efficiency through the expression $I\sim{|{\mbf e}_L\hat R{\mbf e}_S|}^2$, bilinear in the Raman tensor. In the calculations, the Raman tensor does not appear explicitly, the inverse effective mass $\partial^2 E/(\partial k_i\partial k_j)$ playing its role. It is important to note that the Raman efficiency as given by the theory (Eqs.~(\ref{rameff}), (\ref{fdt}), and~(\ref{chir})) is bilinear in the inverse effective mass of the Raman vertex (including the screening part!), i.e. contains the same interferences as the approach involving the Raman tensor. Note that the Tsuneto function $\lambda$ is fully symmetric. In the normal phase, the scattering kernel $\nu$ has been assumed to be the same for all scattering channels. In most of the measurements of the Raman efficiency in orthorhombic high-$T_c$ superconductors, an $A_{1g}$ component has been given. Strictly, this irreducible representation does not exist in $D_{2h}$ but only in $D_{4h}$. In orthorhombic crystals, the Raman tensor contains two $A_g$ components which correspond to the $A_{1g}$ and $B_{1g}$ components of the tetragonal $D_{4h}$ case, and which are not distinguishable in $D_{2h}$ because they transform in the same way. Nevertheless, quantities can be constructed in the orthorhombic case which correspond to the tetragonal $A_{1g}$ component. One of these is $I^{(1)}=(I_{xx}+I_{yy})/2-I_{x'y'}$. Both, $I_{xx}$ and $I_{yy}$ contain $A_{1g}$ and $B_{1g}\,(D_{4h})$, and also an interference term which cancels when $I_{xx}$ and $I_{yy}$ are added. The $I_{x'y'}$ efficiency contains $B_{1g}$ and $A_{2g}\,(D_{4h})$. If we assume that the antisymmetric component ($A_{2g}$ in $D_{4h}$) of the Raman tensor $\hat R$ vanishes (i.e.\ $I_{xy}=I_{yx}$), $I_{x'y'}$ corresponds to tetragonal $B_{1g}$ and cancels the $B_{1g}$ contribution in $I_{xx}$ and $I_{yy}$. Provided that the $A_{2g}$ component of the Raman tensor vanishes, $I^{(1)}$ corresponds to the $I_{A_{1g}}$ of the tetragonal case. Note that the antisymmetric compoment $(R_{xy}-R_{yx})/2$ of the Raman tensor vanishes in the effective mass vertex theory given in Sec.~\ref{sec:gtheory} because of $\gamma_{xy}=\gamma_{yx}$ regardless of the symmetry of the crystal, and also in the experiment in the case of tetragonal crystals but not necessarily for orthorhombic crystals. The equality of $I_{xy}$ and $I_{yx}$ in the calculation is an artifact of the theory. A second possible construction for $A_{1g}$ is $I^{(2)}=I_{x'x'}-I_{xy}$. The $I_{x'x'}$ efficiency contains $A_{1g}$ and $B_{2g}$ contributions. The interference term of these two contributions vanishes in the tetragonal as well as the orthorhombic case. Both, $B_{2g}\,(D_{4h})$ and $A_{2g}$ are contained in $I_{xy}$. But if the $A_{2g}$ component of the Raman tensor vanishes, $I^{(2)}$ also corresponds to the $I_{A_{1g}}$ of the tetragonal case. In one of the experimental works\cite{KraCar} a different method to extract the $A_{1g}$ component was used. Both of the expressions for $I^{(1)}$ and $I^{(2)}$ contain contributions of the $A_{2g}\,(D_{4h})$ Raman tensor component. This component may be present in the experiment, but not in the theory, a fact, that has to be kept in mind when comparing the numerical results to the measurements. Note that the Raman efficiencies in $(xy)$ and $(x'y')$ polarization configurations also contain contributions from the antisymmetric part of the Raman tensor. In view of these uncertainties in $A_{1g}$ we mainly focus in the next section on the directly observable components of the Raman tensor. We shall conclude this section by taking up again the question of the validity of the effective mass approximation. In the experiment, this can be checked in two ways. First, via the dependence of the spectra on the laser frequency which should make it possible to distinguish the contributions to the Raman efficiency resulting from resonant and non-resonant transitions, respectively. The second way involves the measurement of the $A_{2g}$ component of the mass. If the effective mass approximation is valid, the Raman vertex should be symmetric ($\gamma_{xy}=\gamma_{yx}$), i.e.\ the $A_{2g}\,(D_{4h})$ component should vanishes. A non-vanishing $A_{2g}$ component of the measured scattering would cast doubts on the appropriateness of the effective mass approximation. \section{Numerical results and discussion} \label{sec:numerics} To carry out the numerical BZ and FS integrations, we employed a tetrahedron approach.\cite{LehTau,JepAnd} The convergence of the integrations was checked by using different meshes. In Figs.~\ref{Y123} and~\ref{Y124}, the results of full BZ integrations for Y-123 and Y-124, respectively, are plotted. The corresponding spectra obtained through FS integrations can be seen in Ref.~\onlinecite{CarStr}. The Bose factor has not been included, hence the results apply to zero temperature. In both figures, the Raman shift is given in units of the gap amplitude $\Delta_0$. Since the calculated scattering efficiencies for BZ integrations, contrary to FS integrations, are not only a function of the reduced frequency but depend also weakly on the value of $\Delta_0$, we took for the calculations $\Delta_0=220\,{\rm cm}^{-1}$. This value of $\Delta_0$ falls in the range of $\Delta_0$'s determined by Raman scattering and other methods. The delta-function peaks in the Tsuneto function have been broadened phenomenologically by introducing a finite imaginary part $\Gamma=0.3\Delta_0$ of the frequency variable $\omega$. Figures~\ref{Y123} and~\ref{Y124} display spectra for each of the polarization configurations $(yy)$, $(x'x')$, $(xx)$, $(x'y')$, and $(xy)$, as well as the symmetry component $A_{1g}\,(D_{4h})$ (defined by $I_{A_{1g}}=I_{x'x'}-I_{xy}$), the unscreened intensities, the screening part $(\ref{scrpart})$, and the total intensities, equal to the difference between unscreened and screening parts. Note that the $(x'y')$ configuration corresponds to the $B_{1g}\,(D_{4h})$ component because of the vanishing of the $A_{2g}$ component in the theory. We discuss first the results for Y-123. The $A_{1g}$ component (in the rest of this section we use tetragonal notation unless explicitly stated) is subject to rather strong screening, however its unscreened part is comparable to that of the $B_{1g}$ component. The relation between the unscreened and the screened (total) spectral weight of the $A_{1g}$ component is about three. Nevertheless, the shapes of the unscreened and the screened parts are the same and, consequently, {\it there is almost no shift in the peak position due to screening} (contrary to the results of Ref.~\onlinecite{DevEin}). The peak is located almost exactly at $2\Delta_0$. Note that there is no antiscreening in the $A_{1g}$ component. The low-energy part of all $A_{1g}$ spectra (screened and unscreened) is linear, as predicted by the theory. As already mentioned, the $(x'y')$ component (equal to the $B_{1g}$ component in the non-resonant case) is almost four times stronger than its screened $A_{1g}$ counterpart. The screening is very small, its nonvanishing being an effect of the distorted tetragonality of the crystal. There is, in this case, a very small amount of antiscreening in the region below $2\Delta_0$. As in the case of the $A_{1g}$ component, the $(x'y')$ component peaks at almost exactly the $2\Delta_0$ frequency shift. The low-frequency part has an $\alpha\omega+\beta\omega^3$ frequency dependence, the linear part arising from the distorted tetragonality, i.e.\ the fact that the $B_{1g}$ mass does not vanish at exactly the same position on the Fermi surface as the gap function does. The efficiency of the peak in the $(xy)$ configuration (equal to the $B_{2g}$ component in the non-resonant case) is also four or five times smaller than that of the $A_{1g}$ peak. The $(xy)$ peak is located at about $1.3\Delta_0$, as expected from the fact that in the neighborhood of the region where the gap is large, the $B_{2g}$ mass vanishes. Consequently, the peak is not as sharp as in the former cases and screening vanishes since these spectra correspond to a nonsymmetric ($B_{1g}$) representation of the orthorhombic group ($D_{2h}$). In the $A_{1g}$ and $(x'y')$ spectra there should be a small peak at about $\omega=2\sqrt{\epsilon_{\rm vH}^2+\Delta_{\rm max}^2}\approx3.9\Delta_0$ due to the van Hove singularity on the $k_x$-axis near the $X$ point. The corresponding structure, however, is very weak, and practically invisible in Fig.~\ref{Y123}. This is not unexpected for a 3D calculation. These peaks appear strongly when 2D calculations are performed through BZ integrations.\cite{BraCar} In general, the efficiencies in Y-124 (Fig.~\ref{Y124}) are about a factor of three less than those for its Y-123 counterpart. Moreover, the screening of the $A_{1g}$ component of Y-123 is much stronger than that of Y-124. This may be, at least in part, due to the additional chain band: The $(yy)$ component of Y-124 is less screened than the $(yy)$ component of Y-123. At low frequencies, we correspondingly have antiscreening even in $A_{1g}$, a fact which reveals a change of sign of the effective mass on the Fermi surface (see Sec.~\ref{sec:stheory:signch}). Due to this antiscreening, the peak in the $A_{1g}$ spectrum is shifted from $2\Delta_0$ towards approximately $1.6\Delta_0$. In contrast to the situation in Y-123, the Y-124 spectra show clearly the influence of the van Hove singularity on the spectra, as a small hump (vH) located near $2\sqrt{\epsilon_{\rm vH}^2+\Delta_{\rm max}^2}\approx7\Delta_0$. In the $A_{1g}$ spectrum this hump is almost screened out whereas in the $(x'y')$ spectrum it appears slightly increased by the influence of antiscreening. To compare these predictions with the experiment let us first focus on the peak positions. The experimental results for Y-123 (Fig.~\ref{Y123exp}, lower part) clearly show that the position of the $(yy)$, $(x'x')$ and $(xx)$ peaks is at about $300\,{\rm cm}^{-1}$, whereas the $(x'y')$ peak is located at $600\,{\rm cm}^{-1}$, i.e.\ at twice the frequency of the former. This fact is in sharp contrast with the calculated spectra and has been at the center of the controversy concerning the topic at hand.\cite{KCComm,DevRepl} It has been suggested by Devereaux {\it et al.}\cite{DevEin,DevRepl} that the $B_{1g}$ component peaks at $2\Delta_0$, and the $A_{1g}$ component becomes shifted down to almost $\Delta_0$ by the screening. This interpretation contradicts our numerical results which clearly suggest that the influence of screening on the position of the $A_{1g}$ mode is usually smaller. The frequency renormalizations of phonons around $T_c$ also seem to contradict the interpretation in Refs.~\onlinecite{DevEin} and~\onlinecite{DevRepl}. It has been shown\cite{FriTho} that lowering the temperature of the sample in the superconducting phase causes the $A_{1g}$ $435\,{\rm cm}^{-1}$ phonon (plane-oxygen, in-phase) to shift up in frequency and the $B_{1g}$ ($D_{4h}$ notation) $340\,{\rm cm}^{-1}$ phonon (plane-oxygen, out-of-phase) to shift down. This, in turn, implies an amplitude of the gap $2\Delta_0$ between $300\,{\rm cm}^{-1}$ and $360\,{\rm cm}^{-1}$ and is consistent with our interpretation of the electronic Raman spectra with the $A_{1g}$ peak at $2\Delta_0$. Note that the $(yy)$, $(x'x')$ and $(xx)$ spectra do {\it not} contain contributions of the $A_{2g}\,(D_{4h})$ antisymmetric component of the Raman tensor while the $(x'y')$ component does. So, the experimental results may suggest that the shift of the position of the $(x'y')$ spectrum with respect to the peak position of the other spectra is due to resonance effects. The $(xy)$ spectrum is also influenced by the $A_{2g}$ component. It is difficult to determine its peak position from Fig.~\ref{Y123}, but it seems to be located at the same position as that of the $(yy)$, $(x'x')$ and $(xx)$ configurations. The calculation predicts it to be located at about $1.3\Delta_0$, the shift to $2\Delta_0$ can also be attributed to the existence of an $A_{2g}$ component, like in the case of the $(x'y')$ configuration. To compare the relative intensities of the spectra with different polarizations, we refer to Table~\ref{peakh}, which lists them together with the corresponding absolute intensities, both at the peak position. The detailed results of our FS integration have already been reported earlier.\cite{CarStr} We begin with Y-123 (upper panel in Table~\ref{peakh}) and compare BZ integration results to the experimental ones. With the possible exception of the $A_{1g}$ component (and the $(x'x')$ component, which is very similar to $A_{1g}$), the agreement is rather good. The deviation of the $A_{1g}$ component may be attributed to screening, which is very sensitive to sign changes and other details of the Raman vertex near the Fermi surface (such as details of the band structure and especially the exact position of the Fermi energy). The second compound, Y-124 (lower panel in Table~\ref{peakh}), also shows reasonable agreement between the results of the BZ integration and the experiment. However, we also have problems with the $A_{1g}$ component, as we did for Y-123. The measured absolute intensities agree particularly well with the calculations in the case of Y-123. With the exception of $A_{1g}$, the discrepancy between theory and experiment is only a factor of two, which can easily be related to the difficulties in measuring absolute scattering cross sections. In the case of Y-124, the discrepancy is a bit larger, but a factor of four can still be considered good. We should also keep in mind that resonances of $\omega_L$ or $\omega_S$ with virtual interband transitions are expected to enhance the simple effective mass Raman vertex, a fact which could also explain why the measured scattering efficiencies are usually larger than the calculated ones. We close the discussion of the numerical results with a remark about the Fermi surface integration. For Y-124, the results of the former correspond rather closely to the results from the BZ integration. The situation is different for Y-123. Here, the $(xx)$ peak height is almost a factor four larger in the FS integration than in the BZ integration. This is likely to result from the close proximity of the van Hove singularity to the FS in the case of Y-123 ($25\,\meV$), as compared to Y-124 ($110\,\meV$). To verify the predictions related to the effect of orthorhombic distortions as discussed in Sec.~\ref{sec:stheory:odist}, we performed a fit of the function $\alpha\omega+\beta\omega^3$ to the low-frequency part of the $B_{1g}$ data for Y-123 reported in Ref.~\onlinecite{KraCar} and Ref.~\onlinecite{Hacetal} as well as for Bi-2212 (taken from Ref.~\onlinecite{Staetal}) and to the results of our numerical calculations for \hbox{Y-123}. The ratios of the cubic vs.\ the linear part (at $\omega=300\,{\rm cm}^{-1}$) of the fit to the low-frequency efficiency are given in Table~\ref{peakh}. Both measurements for Y-123 agree in their large linear part, which should be due mainly to the lack of exact tetragonality and the presence of impurities. The results of the BZ integration show a smaller linear part, because they do not take into account the influence of impurities. Finally, the result for Bi-2212 is completely different from the former results for Y-123. The linear part almost vanishes, in agreement with the preceding discussion. \section{Conclusions} \label{sec:concl} In spite of the striking ability to predict not only general features of the observed spectra but also their peak intensities, our calculations are not able to predict the relative positions of the $A_{1g}$ and $B_{1g}$ peaks. According to Figs.~\ref{Y123} and~\ref{Y124} the $A_{1g}$ spectrum should peak only slightly below $2\Delta_0$ while $B_{1g}$ should peak at $2\Delta_0$. The experimental data of Figs.~\ref{Y123exp} and~\ref{Y124exp}, however, indicate that the $B_{1g}$ spectra peak nearly at twice the frequency of $A_{1g}$. Since the observed $A_{1g}$ peak is considerably sharper than that of $B_{1g}$, we may want to assign the $A_{1g}$ peak to $2\Delta_0$. Our calculations show that it is impossible to reproduce both peak frequencies with a simple gap of the form $\Delta_0\cos2\phi$ where $\phi$ is the direction of the ${\mbf k}$-vector. A reasonable fit was obtained in Ref.~\onlinecite{KraCar} with a two-dimensional FS which did not take into account the chain component and assigned $d$- and $s$-like gaps to the two bonding and antibonding sheets of the FS of the two planes in an {\it ad hoc} way. Within the present 3-dimensional band structure the FS cannot be broken up into bonding and antibonding plane and chain components since such sheets are interconnected at general points of $k$-space. It is nevertheless clear that there is no reason why the gap function should be the same in the various sheets for a given ${\mbf k}$-direction. Thus the remaining discrepancy in the peak positions between theory and experiment could be due to a more complicated $\Delta_{n{\mbf k}}$ than a simple $\Delta_0\cos2\phi$ used here. Another possible source of this discrepancy is scattering through additional excitations of a type not considered here (e.g.\ magnetic excitations) contributing to and broadening the $B_{1g}$ peak. A BCS-like theory, which involves an attractive pairing potential as well as the repulsive Coulomb potential and uses an anisotropic ${d_{x^2-y^2}}$-like gap function in connection with the effective mass approximation used in the calculation of the absolute Raman scattering efficiencies yields result which are in significant agreement with the experimental spectra. One exception, the peak positions of the $A_{1g}$ and the $B_{1g}$ components, remains unexplained. The theory predicts them to be both located near $\omega=2\Delta_0$, but the experiment shows the peak in $B_{1g}$ at almost twice the frequency of the peak in $A_{1g}$. The weak $B_{2g}$ spectrum agrees in intensity and peak position with calculations for a ${d_{x^2-y^2}}$-like gap. The results of other experiments, involving the temperature dependence of phonon frequencies,\cite{FriTho} suggest that the $A_{1g}$ peak position corresponds to the gap amplitude $2\Delta_0$. The shifting of the $B_{1g}$ peak towards higher frequencies may have an origin different from the mass-fluctuation-modified charge-density excitations described in the theoretical part of this paper but could also be due to a multi-sheeted gap function, more complicated than the simple ${d_{x^2-y^2}}$-like $\Delta_0\cos2\phi$ gap assumed in our calculations. The initial variation of the $A_{1g}$ and $B_{1g}$ scattering efficiencies vs.\ $\omega$ are linear as expected for that gap. The $B_{1g}$ symmetry becomes $A_g$ in the presence of the orthorhombic distortion related to the chains. Consequently, the scattering efficiency at low frequency is not proportional to $\omega^3$ but should have a small linear component which is found both in the calculated and the measured spectra. In the corresponding spectrum of Bi-2212, with and orthorhombic distortion along $(x+y)$, the $B_{1g}$ ($D_{4h}$) excitations also have a nonsymmetric $B_{1g}$ ($D_{2h}$) orthorhombic character. Consequently, for small $\omega$ no component linear in $\omega$ is found in the measured spectra. We have performed our calculations using either BZ or FS integration. In the case of Y-124 the spectra so obtained are very similar. For Y-123 quantitative differences appear; they are probably related to the presence of a van Hove singularity close to the FS. These singularities appear as weak structures in the calculated spectra, as expected for a 3D band structure. \acknowledgments We thank Jens Kircher for providing us with the LMTO band structures. One of us (TS) also would like to thank his colleagues at the MPI for numerous discussions on Raman scattering and superconductors. Thanks are specially due to Igor Mazin for a critical reading of the manuscript.
proofpile-arXiv_065-489
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} Among many versions of the scalar-tensor theory of gravitation, the prototype Brans-Dicke (BD) model [\cite{bd}] is unique for the following four assumptions on the scalar field: (i) a nonminimal coupling of the simplest form; (ii) masslessness; (iii) no self-interaction; (iv) no direct matter coupling. Although the model may not be fully realistic, it still seems to deserve further scrutiny as a testing ground of many aspects of wider class of the scalar-tensor theories. The model has been studied, however, mainly as a classical theory. We attempt to take quantum effects due to the matter coupling into account. It has been argued that composition-independence that entails from the assumption (iv) above would be violated as a quantum correction [\cite{yf1}]. We came to realize, however, that the suspected contribution is canceled by other terms arising from regularization;\footnote{More details on this analysis will be reported in another publication.} WEP is in fact a well-protected and robust property of the model beyond the classical level. We discuss in this note another quantum effect which, as it turns out, has serious cosmological consequences, beyond the extent of remedy expected by adjusting the fundamental parameter of the model. For a possible way out we suggest to modify the assumption (iv), making the scalar field almost ``invisible," thus reconciling with the absence of experimental evidences, still playing a cosmological role. We also discuss how the scalar field acquires a nonzero self-mass due to the matter coupling, even having started with the {\em classical} assumption (ii), but likely in an entirely insignificant manner in practice. We start with the basic Lagrangian \begin{equation} {\cal L}=\sqrt{-g}\left( \frac{1}{2} F(\phi) R -\epsilon\frac{1}{2} g^{\mu\nu}\partial_{\mu}\phi \partial_{\nu}\phi +L_{\rm m}\right), \label{bd5_51} \end{equation} with \begin{equation} F(\phi)=\xi\phi^2. \label{bd5_52} \end{equation} We use the unit system of $8\pi G=1.$\footnote{The units of length, time and energy are $8.07\times 10^{-33}$cm, $2.71\times 10^{-43}$ sec, and $2.43\times 10^{18}$ GeV, respectively. Notice also that the present age of the Universe is $\sim 10^{60}$.} The scalar field $\phi$ and the constant $\xi$ are related to the original notation $\varphi$ and $\omega$, respectively, by \begin{equation} \varphi =\frac{1}{2}\xi\phi^2, \quad{\rm and }\quad \omega =\frac{1}{4\xi}. \label{bd5_53} \end{equation} We also allow $\epsilon =\pm 1$, a minimum extension of the original model to avoid an immediate failure. As we see shortly, $\epsilon =-1$ does not necessarily imply a ghost in the final result. As the matter Lagrangian we choose, according to the assumption (iv), \begin{equation} L_{\rm m}= -\overline{\psi} \left( D\hspace{-.65em}/ +m_{0} \right)\psi, \label{bd55_4} \end{equation} where $\psi$ stands for a simplified representative of the (spinor) matter field. In spite of (iv) $\phi$ couples in effect to the matter field {\em in the field equation.} This is inconvenient, however, when we try to apply the conventional technique of quantum field theory to the $\phi\,$-matter coupling. For this reason we apply a conformal transformation such that the nonminimal coupling is eliminated [\cite{dicke}]: \begin{equation} g_{\mu\nu}=\Omega^{-2}g_{*\mu\nu},\quad{\rm with}\quad \Omega =\left(\xi\phi^2\right)^{1/2}. \label{bd55_5} \end{equation} Notice that we should have \begin{equation} \xi >0, \label{bd55_5a} \end{equation} in order to ensure that the sign of the line element remains unchanged. We in fact find that (\ref{bd5_51}) is now put into the form \begin{equation} {\cal L}=\sqrt{-g_{*}}\left( \frac{1}{2} R_{*} -\frac{1}{2} g^{\mu\nu}_{*}\partial_{\mu}\sigma\partial_{\nu}\sigma +L_{*m} \right), \label{bd3-16} \end{equation} expressed in terms of the new metric $g_{*\mu\nu}$ and the field $\psi_{*}=\Omega^{-3/2}\psi$. Also the {\em canonical} scalar field is now $\sigma$ as defined by \begin{equation} \phi(\sigma)=\phi_{0}\;e^{\beta\sigma}, \label{bd5-2} \end{equation} where \begin{equation} \beta =\frac{1}{\sqrt{6+\epsilon\xi^{-1}}} =\sqrt{\frac{4\pi G}{3+2\epsilon\omega}}. \label{bd5-2a} \end{equation} We emphasize that $\sigma$ is not a ghost if \begin{equation} \beta^2 >0, \label{bd5-2b} \end{equation} even if $\epsilon =-1$ [\cite{fn}]; ``mixing" between the scalar field and (spinless part of) the metric field provides sufficient amount of positive contribution to overcome the negative kinetic part. The condition (\ref{bd5-2b}) is equivalent to \begin{equation} \epsilon\xi^{-1} > -6,\quad {\rm or}\quad \epsilon\omega >-\frac{3}{2}. \label{bd5-2c} \end{equation} We now have a direct $\sigma$-matter coupling expressed in terms of the interaction Hamiltonian to which usual perturbation method is readily applied. We point out, however, that EEP, hence WEP as well, remains intact because the deviation from geodesic arising from the transformation (\ref{bd55_5}) is given entirely in terms of $\sigma$ which is independent of any specific properties of individual particles; the fact that any motion is independent of the mass, for example, verified in one conformal frame obviously survives conformal transformations. We call the conformal frames before and after the conformal transformation J frame and E frame, respectively.\footnote{The name J frame is used, following Cho's suggestion [\cite{cho}], after P. Jordan [\cite{jd}] who was the first to discuss the nonminimal coupling. On the other hand, E frame is a reminder that this is a frame in which the standard theory of Einstein is formulated.} \section{Quantum effect} In E frame, in which we hereafter suppress the symbol $*$ for simplicity, we consider one-loop diagrams as shown in Fig. 1, due to the interaction \begin{equation} H'_{1}=m_{0}\xi^{-1/2}\phi^{-1}(\sigma)\overline{\psi}\psi, \label{bd5-1} \end{equation} coming originally from the mass term in J frame. As will be shown, $\sigma$ may keep moving with time, and so does the mass. But $\sigma$ moves so slowly compared with any of the microscopic time-scales, that the mass $m(t)$ {\em at each epoch} will be defined by \begin{equation} m(t)=m_{0}\xi^{-1/2}\phi^{-1}(\sigma(t)). \label{bd5-3} \end{equation} Now consider the 1\hspace{.2em}-$\phi^{-1}$ diagram (a). Its contribution is given by the potential of $\sigma$; \begin{equation} V_{1}(\sigma)=im(\sigma)\int d^4 p \,{\rm Tr}\! \left(\frac{1}{m(\sigma)+ip\hspace{-.4em}/}\right). \label{bd5-4} \end{equation} Our consideration will be restricted to the Universe which is sufficiently late to justify to ignore the effect of temperature and spacetime curvature.\footnote{The temperature will be lower than $\sim $TeV, for example, if $t\mbox{\raisebox{-.3em}{$\stackrel{>}{\sim}$}} 10^{-13}$sec, much earlier than the epoch of nucleosynthesis. Spacetime curvature will be important only for $t\mbox{\raisebox{-.3em}{$\stackrel{<}{\sim}$}} 10^{-28}$sec, corresponding to the temperature of $10^{11}$GeV.} The integral in (\ref{bd5-4}) is quadratically divergent, but is expected to vanish if there is supersymmetry because the fermionic contribution given by (\ref{bd5-4}) is canceled by the same contribution from the bosonic partner. The cancellation would not be complete, however, if supersymmetry is broken at the mass scale \begin{equation} M_{\rm ssb}=rm, \label{bd5-5} \end{equation} where the ratio $r$ of $M_{\rm ssb}$ to $m$, a representative of ordinary particles taken roughly of the order of GeV, would be $\sim 10^3$--$10^4$, which we naturally choose to be a true constant. The result may be given by \begin{equation} V_{1}(\sigma)=C_{1}r^2 m^4 =V_{1}(0)e^{-4\beta\sigma}, \label{bd5-6} \end{equation} where \begin{equation} V_{1}(0)=C_{1}r^2\left( m_{0}\xi^{-1/2}\phi_{0}^{-1} \right)^4, \label{bd5-6a} \end{equation} with $C_{1}$ most likely of the order 1. Its sign, however, may not be known precisely because it depends on the details of supersymmetry breaking. If $C_{1}>0$, (\ref{bd5-6}) gives a positive exponential potential that would drive $\sigma$ toward infinity. We assume this to be the case. \section{Cosmology} We now consider the cosmological equations with a classical potential $V=V_{1}$ which is the only potential \vspace{.3em}due to the assumption (iii): \begin{eqnarray} &&3H^2 =\frac{1}{2} \dot{\sigma}^2+V +\rho, \label{bd5-7}\vspace{.5em}\\ &&\ddot{\sigma}+3H\dot{\sigma}+V'(\sigma)=0,\vspace{.3em}\label{bd5-8}\\ &&\dot{\rho}+4H\rho =0,\label{bd5-9} \end{eqnarray} where $\rho$ is the matter density which we assume, for the moment, to be relativistic. We also ignored possible terms representing the coupling between $\sigma$ and $\rho$. This would be justified for our purposes as long as we consider late epochs during which the coupling is sufficiently weak [\cite{fn}]. With $V=V_{1}$, a set of analytic solutions of (\ref{bd5-7})--(\ref{bd5-9}) are obtained: \begin{eqnarray} a(t)&=& t^{1/2},\label{bd5-10}\\ \sigma(t)&=&\frac{1}{2\beta}\ln t,\label{bd5-11}\\ \rho(t)&=& \frac{3}{4}\left( 1- \frac{1}{4}\beta^{-2} \right)t^{-2}, \label{bd5-12} \end{eqnarray} with \begin{equation} V_{1}(0)=\frac{1}{16\beta^2}. \label{bd5-12a} \end{equation} Notice that the condition $\rho >0$ is met if \begin{equation} \beta^{-2}<4,\quad {\rm or}\quad \epsilon\xi^{-1}<-2,\label{bd5-13} \end{equation} which, combined with (\ref{bd55_5a}), is satisfied only if $\epsilon =-1$. This is the reason why we decided to allow an ``apparent" ghost in J frame. Then (\ref{bd5-13}) translates into \begin{equation} \xi^{-1}>2,\quad{\rm or}\quad \omega > \frac{1}{2}, \label{bd5-13a} \end{equation} a much milder constraint than those derived from the observation. With $\epsilon =-1$, however, (\ref{bd5-2c}) implies \begin{equation} \xi^{-1}< 6,\quad{\rm or}\quad \omega < \frac{3}{2}, \label{bd5-13b} \end{equation} which is ruled out immediately by the solar-system experiments, giving $\omega\mbox{\raisebox{-.3em}{$\stackrel{>}{\sim}$}} 10^3$ [\cite{will}]. We nevertheless continue our analysis as long as theoretical consistency is maintained. From (\ref{bd5-11}) also follows \begin{equation} \phi(t) =\phi_{0}t^{1/2}, \label{bd5-15} \end{equation} where $\phi_{0}$ is determined by identifying (\ref{bd5-6a}) with (\ref{bd5-12a}); $\phi_{0}^4=16 C_{1}r^2 m_{0}^4 \beta^2 \xi^{-2}$, which, if used in (\ref{bd5-3}), yields \begin{equation} m(t)= \frac{1}{2} \left( C_{1}r^2 \beta^2 \right)^{-1/4}t^{-1/2}. \label{bd5-16} \end{equation} It is interesting to notice that the behavior $m(t)\sim t^{-1/2}$ follows simply because the potential should be proportional to $m^{4}$, as shown in (\ref{bd5-6}), and it must decay like $t^{-2}$ because it is part of the energy density appearing on the right-hand side of the 00-component of the Einstein equation. Obviously the assumption $r=\:$const is crucial in the above argument. We could obtain $m=\:$const if $r(t)\sim t^{-1}$, but with a highly unreasonable consequence that $r$ should be as large as $\sim 10^{63}$ at the Planck time. Also the dependence $V\sim m^4$ is common to any diagrams of many $\phi^{-1}$'s, as in (b) and (c) in Fig. 1. Including them results simply in affecting the overall size of the potential. We arrived at (\ref{bd5-16}) in E frame in which the standard technique of quantum field theory can be applied. We should also notice, however, that we use some type of microscopic clocks in most of the measurements. The time unit of atomic clocks, for example, is provided typically by the frequency $\sim m\alpha^4.$ It is also important to recognize that the cosmic time is usually assumed to be measured in the same time unit. If the time unit $\tau(t)$ itself changes with time, the new time $\tilde{t}$ measured in this unit would be defined by \begin{equation} d\tilde{t}=\frac{dt}{\tau(t)}. \label{bd5-20} \end{equation} In conformity with special relativity, the scale factor $a(t)$ in Robertson-Walker cosmology is transformed in the same manner: \begin{equation} \tilde{a}=\frac{a(t)}{\tau(t)}. \label{bd4-3} \end{equation} These two relations can be combined to a conformal transformation \begin{equation} d\tilde{s}^2 =\tau^{-2}ds^2, \quad{\rm or}\quad g_{\mu\nu}=\tau^2 \tilde{g}_{\mu\nu}. \label{bd4-4} \end{equation} In the prototype BD model, there is no mechanism to make $\alpha$ time-dependent, hence $\tau\sim m^{-1}$. Combining this with $m\sim \phi^{-1}$ as derived from (\ref{bd5-3}), and also comparing (\ref{bd4-4}) with (\ref{bd55_5}), we find that (\ref{bd4-4}) implies going back to the original J frame, as it should because it is the frame in which mass $m_{0}$ is taken to be constant. We now try to solve the cosmological equations in J frame, in which we also suppress tildes to simplify the notation. It is also interesting to find that the behavior $V_{1}\sim \phi^{-4}$ as indicated in (\ref{bd5-6}) shows that this potential in E frame can be derived from a cosmological constant in J frame as given by \begin{equation} \Lambda =C_{1}r^2 m_{0}^4. \label{bd8_4} \end{equation} {\em The quantum effect computed in E frame amounts to introducing $\Lambda$ back in the original conformal frame.} The field equations in J frame are given by \begin{eqnarray} 2\varphi G_{\mu\nu}&=& T_{\mu\nu}+T_{\mu\nu}^{\phi}-g_{\mu\nu}\Lambda -2\left( g_{\mu\nu}\mbox{\raisebox{-0.2em}{\large$\Box$}} -\nabla_{\mu}\nabla_{\nu} \right)\varphi, \label{bd8_6}\\ \mbox{\raisebox{-0.2em}{\large$\Box$}}\varphi &=& \beta^2 (T -4\Lambda),\label{bd8_6a}\\ \nabla_{\mu}T^{\mu\nu}&=& 0.\label{bd8_7} \end{eqnarray} Notice also that $T_{\mu\nu}$ is the matter energy-momentum tensor while \begin{equation} T_{\mu\nu}^{\phi}=\epsilon\left( \partial_{\mu}\phi\partial_{\nu}\phi -\frac{1}{2} g_{\mu\nu}\left( \partial \phi \right)^2\right) . \label{bd8_8} \end{equation} Assuming spatially uniform $\phi$, we derive the cosmological equations: \begin{eqnarray} 6\varphi H^2&=& \epsilon\frac{1}{2} \dot{\phi}^2 +\Lambda +\rho -6 H\dot{\varphi},\label{bd8_9z}\\ \ddot{\varphi}+3H\dot{\varphi}&=&4\beta^2 \Lambda,\label{bd8_9}\\ \dot{\rho} +4 H{\rho}&=&0,\label{bd8_10} \end{eqnarray} where we have chosen $T=0$ confining ourselves to the radiation-dominated era even if $\Lambda$ comes from nonzero $m_{0}$. As a heuristic approach, let us choose \begin{equation} H=0. \label{bd8_11} \end{equation} Then (\ref{bd8_10}) leads to \begin{equation} \rho = {\rm const}. \label{bd8_12} \end{equation} Using (\ref{bd8_11}) in (\ref{bd8_9}) we obtain $\ddot{\varphi}= 4\beta^2 \Lambda,$ which allows a solution \begin{equation} \varphi(t)=2 \beta^2 \Lambda t^2 +\varphi_{1}t +\varphi_{0}. \label{bd8_14} \end{equation} Notice that we have chosen $\Lambda >0$ hence $C_{1}>0$ so that $\sigma$ falls off the potential slope toward infinity in E frame. This implies that $\varphi$ increases also in J frame as indicated in (\ref{bd5-2}) if $\beta^2 >0$. This is the very condition, however, which is in contradiction with the observation, as already discussed in connection with (\ref{bd5-13b}). Taking aside this drawback for the moment again, we expect \begin{equation} \phi\approx \sqrt{\frac{4\Lambda}{6\xi -1}}\,t, \label{bd8_15} \end{equation} at sufficiently late times. Using this together with (\ref{bd8_11}) and (\ref{bd8_12}) in (\ref{bd8_9z}) gives \begin{equation} \rho = 3\Lambda \frac{1-2\xi}{6\xi -1}, \label{bd8_15a} \end{equation} which is positive if (\ref{bd5-13a}) and (\ref{bd5-13b}) are obeyed. An example is shown in Fig. 2, in which we see how $H$ approaches zero, much faster than $t^{-1}$; the Universe quickly becomes stationary after alternate occurrences of expansion ($H>0$) and contraction ($H<0$). This together with other similar examples indicate strongly that there is an ``attractor" to which solutions of different initial conditions would approach asymptotically. Fig. 3, which is a 2-dimensional cross section of the 3-dimensional phase space of $\varphi,\;\dot{\varphi}$ and $\rho$, illustrates how different solutions are attracted to a common destination given by (\ref{bd8_15}) and (\ref{bd8_15a}), which represent in fact a curve in the whole phase space as one finds because of the relation $\dot{\phi}^2=\dot{\varphi}^2/(2\xi\varphi)$. A trajectory for a set of initial values proceeds along, spiraling around and coming ever closer to this curve. One might be tempted to compare our solutions with those in Einstein's model with a negative cosmological constant $\Lambda <0$, but of course without $\phi$. This model allows a static solution $\rho = -\Lambda$ and $H=0$, but the Universe would never become stationary in contrast to our solutions; $\rho$ with a sufficiently large initial value decreases toward the minimum $-\Lambda$ but bounces back to increase in a touch-and-go fashion. In this way we come to conclude that the BD model corrected for an important quantum effect should result in a steady state Universe, which is totally unacceptable in view of the success of the standard model in understanding primordial nucleosynthesis. The constant scale factor in J frame may be interpreted also from the analysis in E frame, in which the length unit is provided by $\tau\sim\phi\sim t^{1/2}$ which increases {\em in the same rate} as $a\sim t^{1/2}$ shown in (\ref{bd5-10}) [\cite{nth}]. We also find that $\phi\sim t$ as given by (\ref{bd8_15}) in J frame and $\phi\sim t_{*}^{1/2}$ in (\ref{bd5-15}), in which the symbol $*$ is restored in E frame, are consistent with each other, since $t\sim t_{*}^{1/2}$ is a consequence of the relation (\ref{bd5-20}).\footnote{Apply the replacement, $\tilde{t}\rightarrow t, t\rightarrow t_{*}.$ For the dust-dominated Universe with $a_{*}\sim t_{*}^{2/3}$, we find $a\sim t^{1/6}.$} These observations seem to support (\ref{bd5-9}) which is only approximate unlike its counter part (\ref{bd8_10}) in J frame. On the other hand, one may ask if there is any sensible solution with $a\sim t^{1/2}$ in J frame. In (\ref{bd8_10}) we substitute \begin{equation} H=(1/2)t^{-1}, \label{bd10_5} \end{equation} thus obtaining \begin{equation} \rho =\rho_{0}t^{-2}. \label{bd10_6} \end{equation} Then (\ref{bd8_9}) becomes \begin{equation} \ddot{\varphi}+\frac{3}{2}t^{-1}\dot{\varphi} =4\beta^2 \Lambda, \label{bd10_7} \end{equation} which is solved asymptotically: \begin{equation} \varphi\approx \frac{4}{5}\beta^2 \Lambda t^2, \quad {\rm or}\quad \phi\approx \sqrt{\epsilon\frac{8}{5}\left( 1 -6\beta^2 \right) \Lambda }\,t. \label{bd10_8} \end{equation} We then find that the right-hand side of (\ref{bd8_9z}) is given by \begin{equation} \rho +\frac{3}{5}\left( 3-16\beta^2 \right)\Lambda. \label{bd10_9} \end{equation} Now from (\ref{bd10_5}) and (\ref{bd10_8}), the left-hand side of (\ref{bd8_9z}) should be time-independent. This can be matched with the situation in which $\rho$ given by (\ref{bd10_6}) decreases rapidly to be negligible compared with the second term of (\ref{bd10_9}); implying that the Universe becomes asymptotically ``vacuum dominated,'' again an unrealistic conclusion. Even worse, ignoring $\rho$ in the right-hand side of (\ref{bd10_9}) and using (\ref{bd10_8}) on the left-hand side of (\ref{bd8_9z}) yields \begin{equation} \beta^2 =\frac{1}{6}, \label{bd10_10} \end{equation} which on substituting into (\ref{bd5-2a}) gives \begin{equation} \xi^{-1}=0, \label{bd10_11} \end{equation} hardly a realistic result.\footnote{The same analysis applied to the dust matter results in $\epsilon =-1$ and $\xi =-2/3$, being inconsistent with (\ref{bd55_5a}).} \section{Discussions} We add that our argument of choosing J frame is independent of whether we literally use atomic clocks to measure something during the epoch in question. It is simply in accordance with realistic situations that analyses are based on quantum mechanics in which mass of every particle is taken to be truly constant. We admit that there should be some other quantum effects to be included. The result obtained here is, however, so remote from what would be expected from the standard scenario that it is highly unlikely that those ``other" effects conspire miraculously to restore the success in the nucleosynthesis, among other things. It seems that we need some more fundamental modification of the model. A possible way out is to abandon the assumption (iv) about the absence of the {\em direct} $\phi$ coupling to the matter in J frame. As an extreme counter example, we may replace the matter Lagrangian (\ref{bd55_4}) in J frame by \begin{equation} L_{\psi}=-\overline{\psi}\left( D\hspace{-.65em}/ +f\phi \right)\psi, \label{bd5-33} \end{equation} where $f$ is a dimensionless Yukawa coupling constant. $\psi$ is massless in J frame, while in E frame we obtain the mass $m= f\xi^{-1/2}$ (in units of $(8\pi G)^{-1/2}$) which is {\em independent of} $\sigma$. The scalar field is {\em decoupled} from $\psi$ {\em in E frame}, hence is left {\em invisible} through the matter coupling; it plays a role {\em only in cosmology}, most likely as a form of dark matter. With {\em constant} mass in E frame, which is now physical, we may reasonably adjust parameters such that the standard scenario is reproduced to a good approximation. There might be intermediate choices between the prototype model and this extreme model. Then we may expect the matter coupling generically {\em much weaker} than \begin{equation} H'_{\sigma_{1}}=-\beta m(t)\overline{\psi}\psi\sigma, \label{bd5-30} \end{equation} in the prototype model, hence evading immediate conflicts with the test of WEP and the constraint from the solar-system experiments.\footnote{See Ref. [\cite{fn}] for a model of this type.} Needless to say, $G$ is predicted to be constant, by construction. With modifications of this nature in mind, we add a comment on the mass term of the scalar field, which is ought to arise from $V_{1}$ as given by (\ref{bd5-6}). We should be interested here in a {\em fluctuating component} $\sigma_{1}(x)$ which is responsible for the force between local mass distributions, to be separated from the spatially uniform component $\sigma_{0}(t)$ evolving as the cosmic time $t$; \begin{equation} \sigma(x)=\sigma_{0}(t)+\sigma_{1}(x), \label{bd5-24} \end{equation} which satisfies \begin{equation} \mbox{\raisebox{-0.2em}{\large$\Box$}}\sigma -V'(\sigma)=0, \label{bd5-25} \end{equation} from which (\ref{bd5-8}) derives. The cosmological component $\sigma_{0}$ is a solution of (\ref{bd5-25}) with $\sigma_{1}$ dropped; \begin{equation} \mbox{\raisebox{-0.2em}{\large$\Box$}}\sigma_{0}+4\beta V_{1}(0)e^{-4\beta\sigma_{0}}=0. \label{bd5-26} \end{equation} If we use this in (\ref{bd5-25}) for the entire field (\ref{bd5-24}), we obtain \begin{equation} \mbox{\raisebox{-0.2em}{\large$\Box$}}\sigma_{1}-4\beta V_{1}(0)e^{-4\beta\sigma_{0}} \left( 1-e^{-4\beta\sigma_{1}} \right)=0. \label{bd5-27} \end{equation} Expanding the terms in the last parenthesis, we find \begin{equation} \mbox{\raisebox{-0.2em}{\large$\Box$}}\sigma_{1}=\mu^2\sigma_{1} +\cdots, \label{bd5-28} \end{equation} where, using (\ref{bd5-11}) for $\sigma_{0}(t)$,\footnote{For dust-dominated Universe, the right-hand side is doubled.} \begin{equation} \mu^2 =16\beta^2 V_{1}(0)t^{-2}, \label{bd5-29} \end{equation} which is $V''(\sigma)$ at $\sigma = \sigma_{0}$. This shows that the scalar field does acquire a ``mass" even though the potential has {\em no stationary point}, but the range of the force mediated by $\sigma_{1}$ is basically given by $t$, which is the size of the visible part of the Universe at each epoch. The force-range at the present epoch can be as ``short" as $10^5$ ly if $16\beta^2 V_{1}(0)\sim 10^{10}$, in contrast to (\ref{bd5-12a}). It is rather likely that the force can be considered to be infinite-range in any practical use. We may relax the assumption $F\sim \phi^2$ in the prototype model. We recognize, however, that the relation of the type $V(\sigma)\sim m^4(\sigma)$ as in (\ref{bd5-6}) is quite generic and so is $m\sim t^{-1/2}$ according to the argument following (\ref{bd5-16}). This makes the conclusion (\ref{bd8_11}) almost inevitable, as long as the assumption (iv) is maintained. As another aspect of more general $F(\phi)$, we point out that the factor $\xi^{-1/2}\phi^{-1}$ in (\ref{bd5-1}) is in fact $F^{-1/2}$. It then follows that the potential as given by (\ref{bd5-6}) generalizes to \begin{equation} V_{1}(\sigma)\sim m^{4}\sim F^{-2}. \label{bd5-34} \end{equation} The relation (\ref{bd5-2}) is also traced back to \begin{equation} \frac{d\sigma}{d\phi}= F^{-1}\sqrt{\epsilon F +\frac{3}{2}F'\:^2}. \label{bd5-35} \end{equation} We may then expect that the potential as a function of $\sigma$ would have a minimum if $F(\phi)$ has a maximum, the same conclusion as in Refs. [\cite{dm}]. The potential minimum should act, however, as an effective cosmological constant, which might present another serious conflict with observations unless it remains below the level of $\sim t^{-2}_{0}\sim 10^{-120}$ at the present epoch. In this respect we have an advantage in the potential having no stationary point. As the last comment we point out that the occurrence of a ghost which was required to give positive matter density is not entirely unnatural from the point of view of unified theories. Consider, for example, Kaluza-Klein approach to $4+n$ dimensional spacetime. The size of compactified {$n$}-dimensional space behaves as a 4-dimensional scalar field, which is shown to have {\em wrong sign} in the kinetic term. This model provides also one of the natural origins of the nonminimal coupling. \begin{center} {\Large\bf Acknowledgments} \end{center} I thank Akira Tomimatsu and Kei-ichi Maeda for valuable discussions. \begin{center} {\Large\bf References} \end{center} \begin{enumerate} \item\label{bd}C. Brans and R.H. Dicke, Phys. Rev. {\bf 124}(1961)925.\item\label{yf1}Y. Fujii, Mod. Phys. Lett. {\bf A9}(1994)3685. \item\label{dicke}R.H. Dicke, Phys. Rev. {\bf 125}(1962)2163. \item\label{fn}Y. Fujii and T. Nishioka, Phys. Rev. {\bf D42}(1990)361. \item\label{cho}Y.M. Cho, Phys. Rev. Lett. {\bf 68}(1992)3133. \item\label{jd}P. Jordan, {\sl Schwerkraft und Weltalle}, (Friedrich Vieweg und Sohn, Braunschweig, 1955).\item\label{will}See, for example, C. Will, {\sl Theory and Experiment in Gravitational Physics,} rev. ed., Cambridge University Press, Cambridge,1993. \item\label{nth}T. Nishioka, Thesis, University of Tokyo, 1991. \item\label{dm}T. Damour and K. Nordtvedt, Phys. Rev. Lett. {\bf 70}(1993)2217; T. Damour and A.M. Polyakov, Nucl. Phys. {\bf B423}(1994)532. \end{enumerate} \newpage \epsfverbosetrue \begin{figure}[h] \hspace*{1cm} \epsfxsize=12cm \epsfbox{abc4.eps} \caption{Some of the one-$\psi$-loop diagrams, shown only up to the third order in $\phi^{-1}$. Each blob represents $\phi^{-1}$ appearing in (12), a collection of many vertices of $\sigma$'s, while a solid line is for the $\psi$ line.} \end{figure} \begin{figure}[h] \centering \hspace*{1cm} \epsfxsize=12cm \epsfbox{bdd2f6c.eps} \caption{A solution of (38)-(40) with the initial values $\varphi_{0}=0.25.\; \dot{\varphi}_{0}=0.0,\; \rho_{0}=0.1$ at $\ln t=1$. We chose $\Lambda = 0.5$ and $\xi =0.4$.} \end{figure} \begin{figure}[h] \hspace{2cm} \epsfxsize=10cm \epsfbox{bdd2f5cc.eps} \caption{Trajectories ($\ln t=1-30$) of the solutions of different initial values $\dot{\varphi}_{0}$ and $\rho_{0}$, as shown in the parentheses, with other parameters the same as in Fig. 2. The point of convergence ($\dot{\phi}^2=1.42857, \rho =0.21429$) is given by (44) and (45). The dotted line is for $H=0$; its left-side for $H>0$.} \end{figure} \end{document}
proofpile-arXiv_065-490
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} Recent advances in the physics of single-electron charging of macroscopic conductors (for general reviews see, e.g., Refs. \onlinecite{mes,sct}) have led to proposals for several new analog and digital electronic devices. Such devices are considered, in particular, to be the most likely candidates to replace silicon transistors in future ultra-dense electronic circuits -- see, e.g., Refs.~\onlinecite{pasct,sasha}. Single-electronics is presently one of the most active areas of solid state physics and electronics, with hundreds of experimental and theoretical works being published annually. We are not aware, however, of any previous attempts to quantitatively compare experimental data for a particular device with results of theoretical analysis including geometrical modeling\cite{knoll}. Such a comparison was the main objective of this work. To that end, we selected one of the simplest devices, the single-electron trap\cite{pasct,fulton}. Figure~\ref{schematic} shows the schematic layout of the circuit we discuss in this paper, which consists of a trap coupled to a single-electron electrometer. We will distinguish two types of conductors (``nodes'') in the circuit: {\em externals}, wires which extend to the edges of the chip and connect to the external measuring devices; and {\em islands}, small metallic segments that are connected each other and to the externals by tunnel junctions. \begin{figure} \begin{center} \leavevmode \epsfxsize=6cm \epsffile{schematic.eps} \end{center} \caption{Schematic view of the 8-junction single-electron trap/electrometer circuit. Islands 12 and 13 are strongly coupled and together provide the energy well for an extra electron. Islands 6-11 form the array separating the well from the drive external 1.} \label{schematic} \end{figure} The trap consists of a larger island, providing the potential well for the extra electron, separated from a voltage-biased ``drive'' external by an $N$-island array. The islands of the array are linked by $(N-1)$ tunnel junctions with low capacitance $C_j$ and tunnel conductance $G_j$: \begin{eqnarray} C_j&\ll&\frac{e^2}{k_B T}\,, \label{lowc} \\ G_j&\ll&\frac{e^2}{h}\,. \label{lowg} \end{eqnarray} \noindent Under condition (\ref{lowg}), each electron is localized inside a single island at any given time. As Fig.~\ref{profile}a shows, the array creates an electrostatic energy barrier $\Delta W\sim~e^2/C_j$ between the drive electrode and the trap island. To inject an additional electron into the trap, a bias voltage $V=V_1-V_2$ is applied to the device. At a certain value $V=V_{+}$ the energy barrier is suppressed: an electron tunnels from the drive external through the array and into the trap island. To extract the electron from the trap island, a voltage $V=V_{-}$ is applied, causing a hole to tunnel from the drive external to the trap island, annihilating the trapped electron. \begin{figure} \begin{center} \leavevmode \epsfxsize=8cm \epsffile{profile.eps} \end{center} \caption{Energy of a typical circuit calculated for three values of $V_1$ ($V_+$, $V_{eq}$, and $V_-$), with $V_2$ = 8 mV: (a) as a function of the position of one electron in the array; (b) as a function of the position of one hole in the array, with one electron in the trap (node 13).} \label{profile} \end{figure} Electrons can also overcome the energy barrier by thermal activation and by macroscopic quantum tunneling of charge (``cotunneling''). At sufficiently low temperatures (\ref{lowc}), the rate of thermally activated hopping over the barrier is roughly~\cite{mes,pasct} \begin{equation} \label{gammat}\Gamma_T \sim \frac{G_j}{C_j} \exp\left(\frac{-\Delta W}{k_B T}\right)\,, \end{equation} \noindent while the rate of spontaneous cotunneling through the barrier scales as \cite{mqt} \begin{equation} \label{gammaq}\Gamma _Q\sim \frac{G_j}{C_j}\left(\frac{G_jh}{4\pi ^2e^2}\right) ^{N-2}\,. \end{equation} \noindent If conditions (\ref{lowc}) and (\ref{lowg}) are satisfied, and the number $N$ of junctions in the array is large enough, the rates of thermal activation and cotunneling may be very low. Thus, the lifetime $\tau_L = (\Gamma_T+\Gamma_Q)^{-1}$ of both the zero-electron and the one-electron states of the trap may be quite long, and the device may be considered bistable. When the voltage $V$ is driven beyond the threshold $V_+$ or $V_-$, the electron or hole tunnels through the array in time $\tau \sim~C_j/G_j$, which may be many orders of magnitude shorter than $\tau _L$. Thus, in principle, the trap can serve as a memory cell. Its contents can be read out non-destructively by capacitive coupling of the trap to the single-electron electrometer\cite {mes,sct,pasct} (see Section \ref{results}). Early attempts to trap single electrons were made by Fulton {\it et al.} \cite{fulton}, using systems with two and four Al/AlO$_x$ junctions of area $\sim~100\times 100$ nm$^2$ at temperatures down to 0.3 K. Their results implied trapping times $\tau _L\simeq 1$ sec. Similar experiments by Lafarge {\it et al.}\cite{lafarge} yielded $\tau _L<1$ sec, much shorter than could be anticipated from formulas (\ref{gammat}) and (\ref{gammaq}). A later attempt\cite{nakazato} used a semiconductor (GaAs) structure with a narrow 2DEG channel instead of a well-defined tunnel junction array. A bistability loop was observed, but its size was not clearly quantized, implying that the number of trapped electrons was much larger than one (the authors estimated this number to be 80-100). Finally, Al/AlO$_x$ trap circuits designed and fabricated at Stony Brook\cite {haus,liji} yielded trapping times of over $10^4$ sec (limited only by observation time). The main goal of the present work was to compare the experimental data obtained for these traps with a quantitative theoretical analysis of the circuits. For this purpose, we have constructed a geometrical model of the circuit, calculated the full matrix of self- and mutual capacitances for the conducting nodes in the model, and simulated static and dynamic properties of the trap using these capacitances. \section{Fabrication} \label{fabrication} Circuits consisting of two layers of partially overlapping nodes were fabricated using the standard shadow mask technique\cite{dolan,ful-dol}. The process begins with a Si substrate, either stripped of oxide or covered by a layer of SiO$_2$ of thickness $H$=500 nm. The substrate is coated with a PMMA/copolymer double layer mask. The circuit pattern is written onto the mask using a scanning electron microscope. Then the mask is developed, the Al circuit elements are deposited onto the substrate, and the mask is lifted off. The fabrication process is described fully in Ref. \onlinecite{liji}. Here we present the essential details. \subsection{Mask} The circuit layout, consisting of a set of line segments, is first specified in a ``mask file'' (Fig.~\ref{mask-big}). A version of the same file is also used to start the computational modeling process (see Section \ref{geometric}). Wide lines in Fig.~\ref{mask-big} represent the parts of the externals that extend from the trap and electrometer to contact pads at the edge of the chip. The narrow lines extending inward from the wide lines (Fig.~\ref{mask-med}) represent the inner parts of the externals. The short, narrow line segments (Fig.~\ref{layout}a) represent islands. \begin{figure} \begin{center} \leavevmode \epsfxsize=8cm \epsffile{mask-big.eps} \end{center} \caption{Mask for complete chip containing several circuits. Pattern for circuits discussed in this paper is circled.} \label{mask-big} \end{figure} \begin{figure} \begin{center} \leavevmode \epsfxsize=8cm \epsffile{mask-med.eps} \end{center} \caption{Closeup of layout of circuits discussed in this paper, showing 20 $\mu$m cutoff radius used in simulations, and external node numbers. Externals have wide ($W = 1 \mu$m) and narrow ($w \simeq$ 50 nm) sections.} \label{mask-med} \end{figure} \begin{figure} \begin{center} \leavevmode \epsfxsize=8cm \epsffile{layout.eps} \end{center} \caption{(a) Central part of mask. (b) AFM image of central part of fabricated circuit. (c) 3-D outline of central part of geometrical model used for capacitance calculation.} \label{layout} \end{figure} The pattern of lines is written on the mask using the electron beam. Upon chemical development, each line in the PMMA becomes a window opening into a larger cavity in the copolymer, which is more susceptible to the electrons. This procedure results in a mask, shown schematically in Fig.~\ref{shadow}, with several suspended bridges. \begin{figure} \begin{center} \leavevmode \epsfxsize=8cm \epsffile{shadow.eps} \end{center} \caption{Schematic view of shadow mask fabrication: (a) first-layer evaporation, (b) oxidation (oxide thickness exaggerated for illustration purposes), (c) second layer evaporation.} \label{shadow} \end{figure} \subsection{Deposition} The aluminum islands are deposited in two layers. The first layer is resistively evaporated in high vacuum directly onto the room-temperature substrate (Fig.~\ref{shadow}a). This layer is then oxidized at $\sim$~10 mTorr O$_2$ for $\sim$~10 min (Fig.~\ref{shadow}b), covering the Al islands with a $\sim~1$ nm layer of AlO$_x$. Before depositing the second layer, the chamber is re-evacuated and the substrate is tilted relative to the Al source. The tilt creates a shift $s$ between these two groups of islands, so that they partially overlap (Fig.~\ref{shadow}c). The AlO$_x$ creates tunnel barriers between the first and second layer islands. In our circuits, the shift $s$ was about 120 nm along the vertical direction in Fig.~\ref{layout}a,b. The second aluminum layer is made thicker than the first, to allow reliable step coverage. An AFM image of the resulting circuit is shown in Fig.~\ref{layout}b. This image exaggerates the island widths because of the finite angle of the AFM tip. Other observations (including SEM imaging) show that the islands oriented perpendicular to the direction of the shift were in fact spatially separated, in the successful samples. Figure~\ref{layout}c shows a simplified model of the central part of the circuit, with externals and islands numbered. There are two islands for each corresponding window in the mask. For example, islands 6 and 7 are the first- and second-layer products of the same window (see also Fig.~\ref{model}c.). Since the two layers of each external overlap each other extensively and are connected to the same voltage/current source, they effectively serve as one conductor. Thus, there is only one external for each corresponding window in the mask. \section{Geometrical Modeling} \label{geometric} The essential electrostatics of a group of conductors can be described by their mutual capacitance matrix, {\boldmath $C$}. A program known as FastCap\cite{fastcap} can calculate {\boldmath $C$} for an arbitrary collection of conductors, given the geometry of the conductors as input. The conductor surfaces are presented to FastCap as a set of discrete elements, or ``panels''. We wrote a program called Conpan (for {\it conductor panels}) to generate a 3D paneling of a simplified model of the experimental system, starting from a 2D mask file. We will first explain the Conpan algorithm, then how its input parameters were derived from experiments. \subsection{Conpan Algorithm} Conpan represents circuit nodes by means of data structures called ``sections''. Each section is a collection of data about a node or part of a node. The data include parameters such as node number, layer number, and limits in the $xy$ plane. Sections may be recursively divided into subsections to represent overlaps and to facilitate paneling. Consider two line segments from the larger mask file (Fig.~\ref{model}a). These two segments eventually produce four islands separated by three tunnel junctions. Conpan expands each segment into a first-layer section (Fig.~\ref{model}b) using the line-width $w$. The second-layer sections (Fig.~\ref{model}c) are initially identical to the first-layer sections except for a uniform translation $s$ that results in overlaps. \begin{figure} \begin{center} \leavevmode \epsfxsize=8cm \epsffile{model.eps} \end{center} \caption{Geometrical model construction in Conpan: (a) Two segments of the array. (b) First-layer sections generated by Conpan for the two segments. (c) Second-layer island sections (shaded), partially overlapping first-layer sections. (d) Shape of first-layer islands. (e) Shape of second-layer islands overlapping first-layer. (f) First- and second-layer islands paneled for capacitance calculation.} \label{model} \end{figure} \subsubsection{Overlap Detection} Since a single second-layer section can overlap more than one first-layer section, Conpan detects the overlaps using a recursive detection algorithm. To begin, each second layer section is compared against each first layer section to detect overlaps. When an overlap is found, the second layer section spawns two daughter sections, one overlapping and one not. The axis and coordinate of the split are stored in the the mother section, along with pointers to the daughter sections. The mother section becomes a placeholder, used only to keep track of the relationship among its daughter sections. The non-overlapping daughter is then is compared against the remaining first-layer sections to find other overlaps. If there are more overlaps, the daughter spawns a pair of sub-daughters, and so on. The recursive process stops when no new overlaps are found. The daughter sections that remain undivided are called ``final daughters''. Figure~\ref{divide} gives a schematic view of the recursive overlap detection process for a single island. \begin{figure} \begin{center} \leavevmode \epsfxsize=8cm \epsffile{divide.eps} \end{center} \caption{Recursive overlap detection in Conpan for a circuit fragment where a second-layer island $S$ overlaps two first-layer islands $F^1$ and $F^2$. (a) Top view, with overlap areas shaded. (b) Comparing $S$ to all first-layer islands, overlap of $F^1$ is detected. $S$ is split into two daughter sections, $S_a$ and $S_b$, at $y = y^*$. (c) Comparing $S_b$ to all first-layer islands, overlap of $F^2$ is detected. $S_b$ is split into $S_{ba}$ and $S_{bb}$ at $y = y^*_b$. (d) Daughter sections $S_a$, $S_{ba}$, and $S_{bb}$, having no further overlaps, are used to build the 3D structure shown in (e).} \label{divide} \end{figure} In addition to dividing up the second-layer sections to account for overlaps, Conpan also splits first-layer sections along the line where they are overlapped (see Fig.~\ref{model}f). This allows the edges of panels facing each other across a junction to line up, facilitating convergence in capacitance calculations. \subsubsection{3D Representation} Once all the overlaps have been found, Conpan can begin to create the 3-D model of the circuit. Each final daughter section becomes the base of a ``block'', a rectilinear solid representing part of a conductor. The heights of the two layers are specified by the parameters $h_1$ and $h_2$. First-layer blocks and non-overlapping second-layer blocks have their base at $z = 0$. Overlapping second-layer blocks have their base at $z = h_1 + t$, where $t$ is the thickness of the gap between overlapping islands that represents the tunnel junction. Finally, each block surface is divided into panels. The goal is to divide the surfaces in such a way that an acceptably accurate capacitance calculation can be performed, within the limits of available computer memory and calculation time. The division process is guided by an input parameter $a$, the goal panel length. The surface of a block with length $L_i$ along axis $i$ is divided into the number $n_i$ of divisions that brings $L_i/n_i$ closest to $a$. Once the block surface has been divided along both its axes, the resulting grid of panels is written to a panel file for input to FastCap. Each panel is stored simply as a quartet of ${x,y,z}$ coordinates, one for each of the four corners, together with the number of the node it belongs to. \subsection{Conpan Input Parameters} \subsubsection{Junction Thickness} In the physical circuit, the tunnel barriers separating the islands consist of AlO$_x$, with unknown $x$ and thickness $t_j$. From literature data on similar junctions \cite{maezawa}, we expect a dielectric constant $\epsilon _j \sim~4$ and $t_j \sim~1$ nm. FastCap can handle dielectric surfaces much as it handles conductors -- by dividing them into panels. However, each additional dielectric panel demands more computer memory and calculation time. Since $t_j$ is much smaller than the transversal dimensions in all junctions, the electric field configuration outside the junctions does not depend strongly on their internal geometry. Therefore, we avoided modeling the junction dielectrics explicitly by replacing them with uniform free-space gaps ($\epsilon =1$) with the effective thickness $t=t_j/\epsilon _j$. This effective thickness was adjusted to make the junction specific capacitance match the standard experimental value $4.5\mu$F/cm$^2$ typical for the Al/AlO$_x$/Al junctions with tunnel conductivity in our range ($\sim~10^5$ S/cm$^2$)\cite{maezawa,mager}. \subsubsection{Line Widths} The effective line width $w$ of islands (and of the narrow parts of externals) is difficult to measure directly, because of its small magnitude (see Fig.~\ref{layout}b). We determined $w$ by requiring that the simulated inverse self-capacitance of the electrometer island ($C^{-1}_{14,14}$) match its experimentally measured value. We derive $C^{-1}_{14,14}$ from the maximum value of the electrometer Coulomb blockade threshold voltage $U_t$, as seen in electrometer I-V plots (Fig.~\ref {elect_i-u}): \begin{equation} \label{c_sigma}(U_t)_{max} = eC^{-1}_{14,14} \simeq e/C_{14,14}\,. \end{equation} \begin{figure} \begin{center} \leavevmode \epsfxsize=8cm \epsffile{elect_i-u.eps} \end{center} \caption{Current-voltage traces for the electrometer, (SiO$_2$/Si substrate, T=35mK, superconductivity suppressed by magnetic field), current biased, taken while rapidly varying $V_5$. The Coulomb blockade voltage $U_t$ varies with $V_5$ from 0 to a maximum value of e$C^{-1}_{14,14}$.} \label{elect_i-u} \end{figure} $C_{14,14}$ (conventionally known as $C_\Sigma$) in turn depends on $w$ because it is is dominated by the electrometer junction capacitances, which increase monotonically with $w$. The experimentally measured values for $C^{-1}_{14,14}$ were $(4.6\times 10^{-16}{\rm F})^{-1}$ for the circuit on Si (sample \#LJS011494B) and $(2.7\times 10^{-16}{\rm F})^{-1}$ for the circuit on SiO$_2$/Si substrate (sample \#LJS011494A). The island width $w$ used in simulation, as determined from Eq. (\ref{c_sigma}), was 30 nm for Si and 42 nm for SiO$_2$/Si. Both of these values are consistent with the values expected from fabrication parameters and from AFM and SEM imaging of the samples. $W$, the width of the wide parts of the externals, is specified as 1 $\mu$m in our mask files, and can be accepted at ``face value'' because it is large compared to the scale of geometrical uncertainty in the circuit, and because the wide parts of the externals are all far (several $\mu$m) from the islands. \subsubsection{Layer Heights} The heights of the two layers ($h_1$ and $h_2$) are determined with a quartz monitor in the deposition unit during fabrication. In our case, these heights were measured to be 30 and 50 nm ($\pm$ 10\%), respectively. \subsection{Substrate} To calculate the effects of the substrate on circuit capacitances, FastCap requires a paneling of the complementary image of the ``footprint'' of the nodes, because panels representing the dielectric/metal interface (parts of the substrate covered by nodes) must be treated differently than panels representing the dielectric/air interface (the exposed substrate). In a manner analogous to that used for conductor panels (see Section~\ref{matrices}), one could investigate various methods of paneling the complementary substrate image in order to minimize the number of panels, while yet retaining an acceptable level of accuracy. Such a paneling algorithm itself is not simple to create. We avoided this problem through an old calculational trick in electrostatics -- the image method. A modified version of FastCap was created, called ImageCap, which can simulate the effects of a single- or double-layer substrate by creating a set (or multiple sets, in the double-layer case) of image panels. \begin{figure} \begin{center} \leavevmode \epsfxsize=8cm \epsffile{image.eps} \end{center} \caption{Image panels in ImageCap: (a) single substrate, (b) double substrate.} \label{image} \end{figure} In the single substrate case, each image panel is formed by reflecting the original panel about the plane of the surface of the substrate (Fig.~\ref{image}a). For the purposes of calculating the electrostatic potential above the substrate, the charge on the image panel is \begin{equation} \label{image1q}q_1=-\frac{\epsilon -1}{\epsilon +1}q\,, \end{equation} where $q$ is the charge on the original panel, and $\epsilon $ is the relative dielectric constant of the substrate. For a substrate covered by an oxide of thickness $H$, an infinite series of image charges is required for an exact representation of the electrostatic effect of the substrate (Fig.~\ref{image}b). However, the distance from the original charge to each successive image charge increases linearly, \begin{equation} z_{2,i}=-2Ti-d\,,\quad i=1,2... \end{equation} \noindent while the value of each successive image charge decreases exponentially, \begin{equation} q_{2,i}=4\beta \frac{\epsilon _1\epsilon _2}{(\epsilon _1+\epsilon _2)^2} (\alpha \beta )^{(i-1)}q\,,\quad i=1,2... \end{equation} \begin{equation} \alpha =\frac{\epsilon _1-1}{\epsilon _1+1}\,,\quad\beta =\frac{\epsilon _1-\epsilon _2}{\epsilon _1+\epsilon _2}\,. \end{equation} Here $\epsilon _1$ and $\epsilon _2$ are the dielectric constants of the surface oxide layer and the bulk substrate, respectively. For our circuits, we have accepted the table values $\epsilon =12.1$ for the bare Si substrate and $\epsilon _1=4.5$, $\epsilon _2=12.1$ for the SiO$_2$/Si substrate ($\alpha =0.64$, $\beta =-0.46)$. The resulting expression for the double-layer image charges, \begin{equation} q_{2,i}= -0.36 \times (0.29)^{(i-1)} q \,,\quad i=1,2... \end{equation} \noindent shows that $q_{2,4}$ is already down by three orders of magnitude from the original charge. In our calculations, adding image levels beyond $q_{2,4}$ made no difference to the result, within a relative error (of the largest self-capacitances) below $\sim~10^{-4}$ . \section{Capacitance Matrices} \label{matrices} \subsection{Matrix Structure} Using the circuit panels generated by Conpan, ImageCap generates the capacitance matrix for the circuit. ImageCap adds the effect of image panels when calculating potentials, and uses no multipole acceleration; otherwise, its algorithms are the same as in FastCap\cite{fastcap}. First, the inverse capacitance matrix for {\it panels} is calculated and inverted. Each element $\widehat C_{ij}$ in the capacitance matrix for {\it nodes} is then formed by summing all the panel capacitance matrix elements linking nodes $i$ and $j$. The charges and potentials on the nodes are related by \begin{equation} \vec q = \mbox{\boldmath $\widehat C$} \vec \phi\,, \label{mat_eq} \end{equation} \noindent so that $\widehat C_{ij}$ is numerically equal to the amount of charge induced on node $j$ when node $i$ is held at unit potential and all other nodes have zero potential. {\boldmath $\widehat C$} is an $N \times N$ matrix, where $N = N_e + N_i$, and $N_e$ and $N_i$ are the numbers of external nodes and island nodes in the circuit, respectively. Ordering all the external nodes before the island nodes, we can write {\boldmath $\widehat C$} in terms of submatrices: \begin{equation} \mbox{\boldmath $\widehat C$}=\left( \begin{array}{c|c} \bigotimes&\mbox{\boldmath $-\tilde C$}\\ \hline \mbox{\boldmath $-\tilde C^T$}&\mbox{\boldmath $C$}\\ \end{array}\right)\,. \label{quad_mat} \end{equation} \noindent Here {\boldmath $C$} is the symmetric $N_i \times N_i$ matrix of island-island capacitances and {\boldmath $\tilde C$} is the $N_e \times N_i$ matrix of external-island capacitances (with elements defined positive, by convention). External-external capacitances (represented above by the $\bigotimes$) are not needed for our simulations. The matrices calculated for our circuits are shown in Tables I and II. Note the up/down alternation of mutual capacitances along the array for the circuit on Si -- e.g., $\tilde C_{1,i}$, the capacitances linking external node 1 to the islands. For example, although island 7 is closer to external 1 than is island 8, $\tilde C_{1,7} \simeq 0.030 \times 10^{-16}$F is smaller than $\tilde C_{1,8} \simeq 0.045 \times 10^{-16}$F (similarly for islands 9 and 10). In {\boldmath $C$}, we see that $\|C_{6,9}\|$ is smaller than $\|C_{6,10}\|$, etc. This phenomenon reflects the influence of the silicon substrate, which, due to its high dielectric constant ($\epsilon \simeq 12$), links externals to the first-layer islands (which lie flat on the substrate) more strongly than to the second-layer islands (which lie partly on top of the first-layer islands). The capacitances for the circuit on SiO$_2$/Si do not show these oscillations as strongly, as we would expect from the smaller permittivity ($\sim~4.5$) of SiO$_2$. \subsection{Model Accuracy} Our model contains three main simplifications related to computational constraints, each of which introduces error into our capacitance matrix calculations. \subsubsection{Free-space Junctions} As noted above, we calculate {\boldmath $\widehat C$} using free-space junctions of thickness $t$ instead of dielectric junctions of thickness $t_j$. (Although initially this approximation was intended for convenience, it later became a necessity as ImageCap does not handle explicit dielectric panels.) The error involved in this approximation was estimated by using FastCap to model a chain of islands in two ways: with explicit dielectric junctions and with free-space junctions. Results for an 8-island chain, with effective dielectric thickness chosen to make the island self-capacitances in both models the same, indicate that the error involved in this approximation is below 1\% for junction-linked islands, and between 1\% and 4\% for non-junction-linked islands. \subsubsection{Paneling} In calculating capacitances, FastCap/ImageCap assigns a uniform charge distribution to each panel. Hence, its accuracy depends on how well the paneling follows changes in change distribution on the node surfaces. Clearly, the denser the paneling, the better the representation of changes in charge distribution. However, panel density is effectively limited by available computer memory. For example, a FastCap simulation with 5000 panels typically requires more than 128 MB. ImageCap uses even more memory, since it calculates all panel interactions directly. We investigated the dependence of calculated capacitance on paneling density for a simple two cube system (Fig.~\ref{2cube}a). \begin{figure} \begin{center} \leavevmode \epsfxsize=8cm \epsffile{2cube.eps} \end{center} \caption{Two cube system with $L/t=10$: (a) Panelings with 1, 3, and 9 panels/side. (b) Capacitances as a function of panel density ($\vec q = \mbox{\boldmath $C$} \vec \phi$).} \label{2cube} \end{figure} The results (Fig.~\ref{2cube}b) suggest that a non-uniform $3 \times 3$ grid (with a 1/10 ratio of edge panel length to central panel length, reflecting the peak in surface charge near the edges) for the smaller, roughly square-shaped node faces (Fig.~\ref {model}f) is sufficient to calculate capacitances with an error below 10\%. This is essentially how we paneled roughly square-shaped island surfaces. For longer faces (Fig.~\ref{paneling}a) we used a larger number of divisions along their length. In an islands-only test circuit, increasing the total number of panels from $\sim 2000$ (corresponding to the $3 \times 3$ grid for roughly square-shaped surfaces) to 6000 resulted in less than 1\% changes in island-island capacitances. Thus we believe that the total error in island-island capacitances due to finite panel density is perhaps only $\sim$~1\%. \begin{figure} \begin{center} \leavevmode \epsfxsize=7cm \epsffile{paneling.eps} \end{center} \caption{Paneling used in our calculations: (a) Trap islands (nodes 12 and 13). (b) Thin parts of externals. (c) Wide parts of externals.} \label{paneling} \end{figure} To reduce the number of panels in the model, the two layers of an external are fused into one where they overlap. The error involved in this simplification is negligible. In addition, the narrow parts of the externals were divided along their length without edge panels (Fig.~\ref{paneling}b). This simplification was found to cause an error in island-external capacitance of $\sim$~5\% when the island and the external are connected by a junction (Fig.~\ref{c-ext}), and $\sim$~1\% otherwise. Finally, wide parts of the external leads were represented by only their top and bottom surfaces (Fig.~\ref{paneling}c), again to save panels. Since the width to height ratio $W/h \simeq 12$, the error introduced by this simplification is negligible. The top and bottom surfaces are divided according to the $3 \times 3$ type scheme described above for islands. Despite the large size of the resulting panels, the error involved in this simple paneling appears to be $\sim$~1\%. \begin{figure} \begin{center} \leavevmode \epsfxsize=8cm \epsffile{c-ext.eps} \end{center} \caption{Effect of panel density on the calculated island-external capacitance. The test circuit had only islands 14 and 15 and narrow parts of externals.} \label{c-ext} \end{figure} \subsubsection{Lengths of Externals} \label{lengths} The calculated capacitance values depend on the lengths of the external wires used in the model. In general, island-external capacitances increase with external length, at the expense of island stray capacitance (capacitance to a ground at infinity); the self-capacitance of islands does not change appreciably. To measure the error introduced by cutting off the externals at a given length, we have calculated capacitance matrices for test circuits with varying external lengths (Fig.~\ref{c-wire}). These circuits consisted of only one island and only the wide parts of the five externals. As a result, the error induced by cutting off externals in these test circuits should be proportionately larger than the error in the complete circuits. Still, the test circuits indicate that the error involved in cutting of the circuit at a radius of 20 $\mu$m (as in our final versions of the complete circuits) was less than 2\%. \begin{figure} \begin{center} \leavevmode \epsfxsize=8cm \epsffile{c-wire.eps} \end{center} \caption{Effect of lead length cutoff on the calculated island-external capacitance. The test circuit had only island 12 and wide parts of externals.} \label{c-wire} \end{figure} \subsubsection{Total Error} Considering the error caused by the above simplifications in the calculation of {\boldmath $\widehat C$} itself, it seems safe to say that the the combined error for any given calculated capacitance matrix element was less than 10\%. Note that we are not yet considering how well the geometrical model corresponds to the physical circuit (see Section~\ref{discussion}). \section{Simulated and Experimental Results} \label{results} We have calculated most properties of our circuits using {\sc moses}, the single-electron circuit simulation program\cite{moses}. This program uses a Monte Carlo algorithm to simulate arbitrary SET circuits within the framework of the orthodox theory of single-electron tunneling \cite{mes,sct}. {\sc moses} needs to know the capacitance sub-matrices {\boldmath $C$} and {\boldmath $\tilde C$} and the conductances of all tunnel junctions. The resistance of two electrometer junctions connected in series can be extracted from the slope of the experimental dc $I-V$ curve of the electrometer at high voltage ($V\gg e/C_\Sigma$). From this measurement, we calculated tunnel conductance per unit area. Conductances of all other junctions in the circuit were then calculated by assuming that their conductance is proportional to their nominal area. This assumption may only be accurate to an order of magnitude; however, most of the results discussed below pertain to stationary properties of the system, and are thus unaffected by deviations in conductance. \subsection{General Electrostatic Relations} Solving the matrix equation (\ref{mat_eq}) for the island potentials $\phi_i$, with our definition (\ref{quad_mat}) of the capacitance matrix we get \begin{equation} \phi_i=\sum_{j\in isl}{C^{-1}_{ij}(q_j + \tilde q_j)}\,,\quad \tilde q_j \equiv \sum_{k\in ext}{\tilde C_{kj}V_k}\,, \end{equation} \noindent or, in a different form, \begin{equation} \phi_i=\sum_{j\in isl}{C^{-1}_{ij}q_j}+\sum_{k\in ext}{\alpha_{ik}V_k}\,,\quad \alpha_{ik} \equiv \sum_{j\in isl}{C^{-1}_{ij}\tilde C_{kj} }\,, \label{phi_i_def} \end{equation} \noindent where $V_k$ are the external potentials. These relations allow us to establish useful relations between changes in the external potentials $\{V_k\}$ and the charge state of the islands $\{q_i\}$, and the dynamics of the system as determined by the island potentials $\{\phi_i\}$. \subsection{Electrometer} Let us apply these relations, in particular, to the island of the single-electron transistor (number 14 in our notation, see Fig.~\ref{schematic}) serving as the electrometer. Experimentally, we measure the dc voltage $U_{3-4}$ between the ``source'' and ``drain'' of the transistor (externals 3 and 4) under a small ($\sim$~100 pA) dc current bias. If the temperature is small enough ($k_B T \ll e^2C^{-1}_{14,14}$), the voltage $U_{3-4}$ in such an experiment closely follows the threshold $U_t$ of the Coulomb blockade of the transistor -- see Fig.~\ref{elect_i-u}. It is well known (see, e.g., Refs.~\onlinecite{mes,sct}) that the threshold is determined by the effective background charge $Q_o$ of the transistor island, which may be defined as \begin{equation} \phi_{14}|_{U_3=U_4=0} = C^{-1}_{14,14}(q_{14}+Q_o)\,. \label{qo_intro} \end{equation} \noindent Comparing (\ref{qo_intro}) and (\ref{phi_i_def}) above, we obtain in our notation \begin{equation} Q_o = \frac{1}{C^{-1}_{14,14}} \left[ \sum_{j\in isl}^{j\ne 14}{C^{-1}_{14,j}q_j}+\sum_{k\in ext}{\alpha_{ik}V_k} \right]\,. \label{qo_sums} \end{equation} Eq. (\ref{qo_sums}) allows us to find the theoretically expected variation of $Q_o$ due to any changes in the system. On the other hand, the threshold voltage is an e-periodic function of $Q_o$, and its maximum amplitude is expressed by Eq. (\ref{c_sigma}) (for the case when the two transistor junction capacitances are the same). Thus, after we measure the experimental value of $(U_t)_{max}$, we can express the change in the effective charge $Q_o$ via the observed variation in $U_t$: \begin{equation} \Delta Q_o = \frac{e\Delta U_t}{2(U_t)_{max}}\,. \label{qo-ut} \end{equation} We have applied this approach to compare experiment and theory for two samples (\#LJS011494A with SiO$_2$/Si substrate and \#LJS011494B with Si substrate). \subsection{High-$T$ electrometer response} We can readily measure $\Delta V_i$ (i = 1,2,5), the change in external voltage corresponding to one period of the oscillating threshold voltage (Fig.~\ref{elect_u-v5}). At $k_BT\geq 0.1e^2/C_\Sigma$ (experimentally, $T\geq 0.5K$), thermal activation of electrons smears the Coulomb blockade effects and makes the junctions essentially transparent to tunneling, while the periodic response of the electrometer is still visible up to $k_BT \sim 0.3e^2/C_\Sigma$ ($T \sim~1.5$K). Thus, the measured values of $\Delta V_i$ depend only on the circuit geometry and are essentially independent of the properties of the junctions. \begin{figure} \begin{center} \leavevmode \epsfxsize=8cm \epsffile{elect_u-v5.eps} \end{center} \caption{A typical experimental dependence of electrometer dc voltage $U_{3-4}$ on gate voltage $V_5$ for a circuit comprising only an electrometer.} \label{elect_u-v5} \end{figure} {\sc moses} is not a useful tool for directly modeling high-$T$ behavior, as the number of jumps involved would be extremely high. However, we can simulate high-$T$ behavior in {\sc moses} by specifying very high external voltages $V_i$ while keeping temperature low (say, $T=0$). Under these high voltage conditions, the islands are flooded with extra electrons, and the tunnel junctions become effectively transparent to tunneling, just as in the high temperature case. Thus we simply apply an external voltage $V_i \gg e/C_\Sigma$, measure how many electrons enter the electrometer island, and find the ratio of voltage $V_i$ to electrometer charge $q_{14}$. Table III shows values of the ratio $\Delta V_i$ for simulated circuits and for experimental circuits averaged over several nominally identical samples. For the experimental values, the uncertainties given reflect the spread of among the samples. For the simulated values, the uncertainties given reflect the $\sim$~10\% error in calculated values, as described in Section~\ref{matrices}. The simulated values are all lower than the experimental ones (with the exception of $\Delta V_5$ on SiO$_2$/Si), differing by as much as 50\%. The agreement is somewhat better for the circuits on SiO$_2$/Si. \subsection{Trap phase diagram} The simplest measurable characteristic involving single-electron charging of the trap is its phase diagram (Fig.~\ref{phase}), which reflects changes in the charge states of the array and trap as a function of the drive voltage $V_1$. In {\sc moses}, we can directly view the charge state of each island in the array and trap as we vary $V_1$, as well the resulting change in $Q_o$. In the physical circuit, however, we can only measure the response $U_{3-4}$ of the electrometer and reduce it to the changes in $Q_o$ using Eq.~\ref{qo-ut}. \begin{figure} \begin{center} \leavevmode \epsfxsize=8cm \epsffile{phase.eps} \end{center} \caption{Electrometer phase diagram: $Q_0$ as a function of trap drive voltage. $V_2$=8.2mV, sample on SiO$_2$/Si. Crosstalk from drive voltage to electrometer has been subtracted, leaving only influence of trap charge state. In experiment, superconductivity in aluminum is suppressed by a 2T magnetic field.} \label{phase} \end{figure} Figure ~\ref{phase} shows experimental and simulated electrometer response to ramping the trap drive voltage $V_1$ up and down over a period of several minutes. In both cases, the effects of the crosstalk between external 1 and the electrometer have been removed. In the experiment, the crosstalk is cancelled by feeding the electrometer gate (node 5) with a voltage $V_5=-\alpha V_1$, with the coefficient $\alpha$ adjusted to make the phase diagram plateaus horizontal. In simulations, {\sc moses} accomplishes the same effect by subtracting $\Delta \phi_{14} = \alpha_{1,14}V_1$ from the electrometer island potential. Horizontal plateaus in Fig.~\ref{phase} correspond to particular charge states of the system (trap~+~array), while vertical jumps correspond to changes of charge state. Thus, the hysteretic loops are regions of bi/multi-stability. The blow-up of the theoretical curve (Fig.~\ref{loops}) indicates the states for several plateaus. In particular, notice that the largest plateaus correspond to states that are most stable because the array is either charged uniformly (one electron on each island, for example) or in a regular alternating pattern such as 1-0-1-0 (Fig.~\ref{hi-loops}). The smaller plateaus correspond to more complex charge states which are less stable. \begin{figure} \begin{center} \leavevmode \epsfxsize=8cm \epsffile{loops.eps} \end{center} \caption{Closeup view of experimental and simulated phase loops. Charge vectors for islands 6-13 on various plateaus in simulated phase diagram are indicated in brackets. In simulation, $\vec q_o = 0$ was assumed.} \label{loops} \end{figure} \begin{figure} \begin{center} \leavevmode \epsfxsize=8cm \epsffile{hi-loops.eps} \end{center} \caption{Simulated phase loops at a higher range in $V_1$, showing wider plateaus around [1010123].} \label{hi-loops} \end{figure} The experimental phase diagram bears a qualitative resemblance to the theoretical one, with somewhat shorter plateaus, though the order of magnitude is the same ($\sim$~2 mV for major plateaus). Simulated phase diagrams with randomly selected $\vec q_0$ show shorter plateaus than the $\vec q_0 = 0$ phase diagram (see Sec.\ref{discussion} below). In Fig.~\ref{phase}, the large jumps in $Q_o$ correspond to a single electron entering the trap: $\delta Q_o \equiv (\Delta Q_o)|_{e\to tr}$. Using Eq.~(\ref{qo_sums}), we can also express the simulated value of $\delta Q_o$ as \begin{equation} \delta Q_o^{sim} = \frac{C^{-1}_{tr,14}}{C^{-1}_{14,14}}e\,, \end{equation} \noindent where $tr$ = 12 or 13, depending on which trap island the electron stops in. For comparison with experimental results, we take the average of the two possible values. The results are shown in Table IV. The difference between simulated and experimental values for Si is within the estimated geometric calculation error (10\%), while the value for SiO$_2$/Si is not. \subsection{Plateau dependence on $V_2$} For a given plateau, the switching voltages $V_1$ = $V_\pm$ depend on the ``ground'' voltage $V_2$ (see Fig.~\ref{schematic}). In the simplest model, with no stray capacitances, (see, e.g., Ref. \onlinecite {pasct}) the charge state of the system depends only on the voltage $V=V_1-V_2$. In that model, the dependences $V_\pm(V_2)$, corresponding to changes in the charge state, would form parallel 45$^{\circ}$ lines in the $[V_1,V_2]$ plane. In reality, however, stray capacitances of the islands to ``infinity'' (i.e. to a distant common ground) make the average potential $(V_1+V_2)/2$ of the system relevant as well. As a result, the region corresponding to each charge state acquires a shape similar to a stretched diamond (Fig.~\ref{diamond}). \begin{figure} \begin{center} \leavevmode \epsfxsize=8cm \epsffile{diamond.eps} \end{center} \caption{Experimental and simulated threshold voltages $V_\pm$ as functions of $V_2$: Si substrate, superconductivity suppressed.} \label{diamond} \end{figure} Simulations using {\sc moses} show that the diamond shape results from the alternation of two types of electron transport that switch the charge state. At the low-$V_2$ end of the diamond, the charge state switches with the transfer of an electron in/out of the trap (see the energy diagram in Fig.~\ref{profile}a). However, at the high-$V_2$ side, the barrier for holes to enter or exit is lower than for electrons (cf. Fig.~\ref{profile}b). Near the sharp ends of the diamond, the critical transport may be even more complex (e.g., creation of an electron-hole pair inside the array, with the sequential motion of its components apart, one into the trap, and another into the external electrode). In these regions, however, the plateau corresponding to the charge state of the trap is already small and virtually disappears among numerous plateaus corresponding to various internal charge states of the array (Fig.~\ref{loops}). Figure~\ref{diamond} shows that while the diamond shape of the charge state in $[V_1,V_2]$ is well reproduced in experiment, the simulated width \mbox{$\|V_{+}-V_{-}\|$} of the bistability region in $V_1$ is roughly twice the experimental value. For each $V_2$, there is one value of $V_1$, called $V_{eq}$, at which the energy barrier is the same for an electron to tunnel into or out of the trap\cite{seneca}. A good measure of the relative influence of the two external voltages on the trap is the derivative \begin{equation} \frac{dV_{eq}}{dV_2} = \frac{\alpha_{2,tr}}{1 - \alpha_{1,tr}}\,, \end{equation} \noindent where $tr$ = 12 or 13, depending on which trap island actually traps the electron for a given $(V_1,V_2)$. The two values are typically within 5\% of each other, and we take their average when comparing simulated and experimental results. In the experimental data, we define the average $V_{eq}$ by bisecting the diamond shape in the graph (Fig.~\ref{diamond}). $dV_{eq}/dV_2$ is essentially a geometric property of the circuit and should not depend on thermal activation or cotunneling. As Fig.~\ref{diamond} shows, the simulated and experimental values are very close. \subsection{Energy barrier} At $V_1=V_{eq}(V_2)$, we can measure the energy barrier $\Delta W$ experimentally by measuring trapping lifetime as a function of temperature (for experimental details, see Ref.~\onlinecite{haus}). The Arrhenius law for lifetimes gives \begin{equation} \tau_L \propto \exp (\Delta W/kT), \label{arrhen} \end{equation} \noindent so that plotting $\log (\tau_L)$ vs. $1/T$ gives us $\Delta W$. Dynamical simulations\cite{seneca} have shown that (\ref{arrhen}) is virtually unaffected by cotunneling for relatively high temperatures ($\sim~100$ mK and above). In simulation, {\sc moses} allows us to measure $\Delta W$ directly. Figure~\ref{dW_v2} shows the dependence of the trap energy barrier $\Delta W$ on the bias voltage $V_2$, for the circuit on the SiO$_2$/Si substrate. The simulated energy barrier profile peaks at roughly the same value of $V_2$ as in the experiment, and the peak barrier value is within $\sim~10\%$ of the experimental value. However, the simulated peak is sharper. \begin{figure} \begin{center} \leavevmode \epsfxsize=8cm \epsffile{dW_v2.eps} \end{center} \caption{Experimental and simulated dependence of barrier height on $V_2$ (SiO$_2$/Si substrate).} \label{dW_v2} \end{figure} \section{Discussion} \label{discussion} Let us first discuss the results independent of the single-electron charging effects: the transistor response $\delta Q_o$ to a single electron entering the trap, the oscillation periods $\Delta V_i$, and the slope $dV_{eq}/dV_2$. The differences between simulated and experimental values for $\delta Q_o$ are 6\% and 23\% for the Si and SiO$_2$/Si substrates, respectively. Values for $\Delta V_i$ do not agree as well: differences between simulated and experimental values range from 15 to 38\% for the trap on SiO$_2$/Si, and from 29 to 50\% for the trap on Si. We had experimental data for $dV_{eq}/dV_2$ only on Si. Here the difference between experimental and simulated results was $\sim~12\%$. These numbers suggest how well our geometrical model corresponds to the physical circuit (the accuracy of the orthodox theory and of the {\sc moses} simulator is presumably much higher). The most obvious idealization involved in our geometric modeling is that the islands created by Conpan are rectilinear and uniform. Even at the limited resolution of an AFM image (Fig.~\ref{layout}b), the contours of the fabricated circuits appear rounded and irregular on a scale of $\sim$ 10 nm. This is to be expected, due to the relatively large grain size of evaporated Al ($\sim~50$ nm, comparable to the line width $w$) and the stochastic nature of the grain growth process. Most capacitance matrix elements should not depend strongly on small details of the island shape. However, irregularities in the shape of overlapping islands may change the area, and thus the capacitance, of the junctions linking them. All other results involve single-electron charging effects. Here the difference between the theory and experiment is larger - typically by a factor of 2, and sometimes larger. We believe that the most important origin of this difference is the set of background charges $\vec{q}_0$. The Si substrate is capable of trapping charged impurities near the circuit islands. The result of these impurities is that the charge on island $i$ effectively changes from $\tilde{q}_i$ to $\tilde{q}_i + q_{0i}$. These charges may furthermore be capable of thermal migration over time. Simulated plots of the electrometer response to trap charging with three randomly selected $\vec q_0$ are shown in Fig.~\ref{multi_q0}. It appears that the wide ($\sim$~4 mV) steps near $V_1 = 0$ in the $\vec q_0 = 0$ plot are not stable to variations in $\vec q_0$: in most plots with random $\vec q_0$, as in the experimental plot, all step widths are less than 3 mV. \begin{figure} \begin{center} \leavevmode \epsfxsize=8cm \epsffile{multi_q0.eps} \end{center} \caption{Simulated response of electrometer to ramps of $V_1$ for $\vec q_0 = 0$ (solid line) and three randomly selected background charge vectors $\vec q_0$. Plots are shifted vertically because background charge on the electrometer island (node 14) effectively shifts $Q_o$.} \label{multi_q0} \end{figure} To summarize: we have developed an automated way to construct simplified models of experimental single-electron devices and circuits with metallic islands, and we have compared the properties of model single-electron traps with those of real traps. The observed differences between simulation and experiment may be attributed to random deviations of the physical structures from their nominal size and shape, and to random background charges created by charged impurities. Future work of interest may include more precisely defined single-electron devices, using better fabrication technology, and the extension of quantitative modeling to semiconductor-based single-electronic circuits and hybrid single-electronic / conventional transistor logic circuits. As single-electronics evolves into a mature technology, such modeling will be essential. \section{Acknowledgments} We greatly appreciate numerous fruitful discussions with D. Averin, R. Chen, L. Fonseca, A. Korotkov, W. Zheng, and K. Nabors. This work was supported in part by AFOSR grants \#F49620-1-0044 and \#F49620-96-1-0320.
proofpile-arXiv_065-491
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} After the observation of the top quark signal at the Tevatron, the mechanism of the spontaneous electroweak symmetry breaking remains the last untested property of the Standard Model ($\cal{SM}${}). Although the recent global fits to the precision electroweak data from LEP, SLC and Tevatron seem to indicate a preference to a light Higgs boson $m_H=149^{+148}_{-82}$~GeV, $m_H<450$~GeV ($95\%C.L.$) \cite{LEPEWWG}, it is definitely premature to exclude the heavy Higgs scenario. The reason is that a restrictive upper bound for $m_H$ is dominated by the result on $A_{LR}$, which differs significantly from the $\cal{SM}${} predictions \cite{ALR}. Without $A_{LR}$ upper bound on $M_H$ becomes larger than 600~GeV \cite{ALR}, which is not far in the logarithmic scale from the value of the order of 1~TeV, where perturbation theory breaks down. In order to estimate the region of applicability of the perturbation theory, the leading two--loop ${\cal O}(g^4 m_H^4/M_W^4)$ electroweak corrections were under intense study recently. In particular, the high energy weak--boson scattering \cite{scattering}, corrections to the heavy Higgs line shape \cite{GhinculovvanderBij}, corrections to the partial widths of the Higgs boson decay to pairs of fermions \cite{fermi_G,fermi_DKR} and intermediate vector bosons \cite{vector_G,vector_FKKR} at two--loops have been calculated. All these calculations resort at least partly to numerical methods. Even for the two--loop renormalization constants in the Higgs scalar sector of the $\cal{SM}${} \cite{GhinculovvanderBij,MDR} complete analytical expressions are not known. In this paper we present our analytic results for these two--loop renormalization constants, evaluated in the on--mass--shell renormalization scheme in terms of transcendental functions $\zeta(3)$ and the maximal value of the Clausen function $\mbox{Cl}(\pi/3)$. \section{Lagrangian and renormalization} The part of the $\cal{SM}${} Lagrangian describing the Higgs scalar sector of the $\cal{SM}${} in terms of the bare quantities is given by: \begin{eqnarray} &&{\cal L} = \frac{1}{2}\partial_\mu H_0\partial^\mu H_0 + \frac{1}{2}\partial_\mu z_0\partial^\mu z_0 + \partial_\mu w^+_0\partial^\mu w^-_0 \nonumber\\ && -\, \frac{{m_H^2}_0}{2v_0^2}\left(w_0^+w_0^- + \frac{1}{2}z_0^2 + \frac{1}{2}H_0^2 + v_0 H_0 + \frac{1}{2}\delta v^2 \right)^2. \label{lagrangian} \end{eqnarray} Here the tadpole counterterm $\delta v^2$ is chosen in such a way, that a Higgs field vacuum expectation value is equal to $v_0$ \begin{equation} v_0 = \frac{2\, {M_W}_0}{g} \end{equation} to all orders. Renormalized fields are given by \begin{equation} H_0 = \sqrt{Z_H}\,H, \quad z_0 = \sqrt{Z_z}\,z, \quad w_0 = \sqrt{Z_w}\,w. \end{equation} At two--loop approximation the wave function renormalization constants, tadpole and mass counterterms take the form \begin{eqnarray} \sqrt{Z} &=& 1 + \frac{g^2}{16\pi^2}\,\delta Z^{(1)} + \frac{g^4}{(16\pi^2)^2}\,\delta Z^{(2)}; \nonumber \\ \delta v^2 &=& \frac{1}{16\pi^2}\,{\delta v^2}^{(1)} + \frac{g^2}{(16\pi^2)^2}\, {\delta v^2}^{(2)}; \label{constants} \\ {M_W}_{0} &=& M_W + \frac{g^2}{16\pi^2}\, \delta M_W^{(1)} + \frac{g^4}{(16\pi^2)^2}\, \delta M_W^{(2)}; \nonumber \\ {m_H^2}_{0} &=& m_H^2 + \frac{g^2}{16\pi^2}\, {\delta m_H^2}^{(1)} + \frac{g^4}{(16\pi^2)^2}\, {\delta m_H^2}^{(2)}. \nonumber \end{eqnarray} Since weak coupling constant $g$ is not renormalized at the leading order in $m_H^2$, the $W$--boson mass counterterm is related to the Nambu--Goldstone wave function renormalization constant by the Ward identity ${M_W}_0=Z_w M_W$. In the on--mass--shell renormalization scheme all the counterterms are fixed uniquely by the requirement that the pole position of the Higgs and $W$--boson propagators coincide with their physical masses and the residue of the Higgs boson pole is normalized to unity. The one--loop counterterms equivalent to those used in \cite{GhinculovvanderBij,MDR} are given by \begin{eqnarray} {\delta v^2}^{(1)}&=& m_H^{2}{\xi^\epsilon_H}\,\Biggl\{-{\frac {6}{\epsilon}}+3 -\, \epsilon \biggl(\frac{3}{2}+\frac{\pi^2}{8}\biggr) \Biggr\}; \nonumbe \\ \frac{{\delta m_H^2}^{(1)}}{m_H^2}&=& \frac{m_H^{2}{\xi^\epsilon_H}}{M_W^2} \,\Biggl\{-{\frac {3}{\epsilon}}+3-{\frac {3\,\pi\,\sqrt {3}}{8}} +\, \epsilon \biggl(-3+\frac{\pi^2}{16}+\frac{3\pi\sqrt{3}}{8} +\frac{3\sqrt{3}C}{4}-\frac{3\pi\sqrt{3}\log{3}}{16}\biggr) \Biggr\}; \nonumbe \\ \delta Z_H^{(1)}&=&{\frac {m_H^{2}{\xi^\epsilon_H}\,}{M_W^{2}}} \Biggl\{\frac{3}{4}-{\frac {\pi\,\sqrt {3}}{8}} +\, \epsilon \biggl(-\frac{3}{4}+\frac{3\pi\sqrt{3}}{32} +\frac{\sqrt{3}C}{4}-\frac{\pi\sqrt{3}\log{3}}{16}\biggr) \Biggr\}; \label{zh} \\ \frac{\delta M_W^{(1)}}{M_W}&=&\delta Z_w^{(1)}= -{\frac {m_H^{2}{\xi^\epsilon_H}}{16\,M_W^{2}}}\Biggl\{1 -\frac{3}{4}\epsilon\Biggr\}. \nonumbe \end{eqnarray} Here the dimension of space--time is taken to be $d=4+\epsilon$ and \begin{equation} \xi^\epsilon_H=e^{\gamma\,\epsilon/2}\, \left(\frac{m_H}{2\,\pi}\right)^\epsilon. \end{equation} In contrast to papers \cite{GhinculovvanderBij,MDR} we prefer not to keep the one--loop counterterms of ${\cal O}(\epsilon)$ order, {\it i.e.} unlike in the conventional on--mass--shell scheme used in \cite{GhinculovvanderBij,MDR}, we require that the one-loop normalization conditions are fulfilled only in the limit $\epsilon\to 0$, where the counterterms of the order ${\cal O}(\epsilon)$ do not contribute. Such a modified on--mass-shell scheme is equally consistent as the conventional one or the standard scheme of minimal dimensional renormalization, which assumes only the subtraction of pole terms at $\epsilon=0$ \cite{BR}. (Moreover, in general one cannot subtract all the nonsingular ${\cal O}(\epsilon)$ terms in the Laurent expansion in $\epsilon$, as they are not polynomial in external momenta.) The ${\cal O}(\epsilon)$ one--loop counterterms considered in \cite{GhinculovvanderBij,MDR} do can really combine with the $1/\epsilon$ terms at two--loop order to give finite contributions, but these contributions are completely canceled by the additional finite parts of the two--loop counterterms, fixed through the renormalization conditions in the on--mass--shell renormalization scheme. The reason is that after the inclusion of the one--loop counterterms all the subdivergences are canceled and only the overall divergence remains, which is to be canceled by the two--loop counterterms. The account of finite contributions coming from the combination of ${\cal O}(\epsilon)$ one--loop counterterms with $1/\epsilon$ overall divergence just redefines the finite parts of the two--loop counterterms. An obvious advantage of this modified on--mass--shell scheme is that the lower loop counterterms once calculated could be directly used in higher loop calculations, while in the conventional on--mass--shell scheme for $l$-loop calculation one needs to recalculate the one-loop counterterms to include all the terms up to ${\cal O}(\epsilon^{l-1})$, two-loop counterterms to include ${\cal O}(\epsilon^{l-2})$ terms and so on. \section{Analytic integration} The calculation of the Higgs and $W$--boson (or Nambu--Goldstone $w$, $z$ bosons) two--loop self energies, needed to evaluate the renormalization constants (\ref{constants}), reduces to the evaluation of the two--loop massive scalar integrals \begin{eqnarray} &&J(k^2; n_1\, m_1^2,n_2\, m_2^2,n_3\, m_3^2,n_4\, m_4^2,n_5\, m_5^2) = -\frac{1}{\pi^4}\int\,D^{(d)}P\,D^{(d)}Q \,\biggl(P^2-m_1^2\biggr)^{-n_1}\\ && \times \biggl((P+k)^2-m_2^2\biggr)^{-n_2} \biggl((Q+k)^2-m_3^2\biggr)^{-n_3}\biggl(Q^2-m_4^2\biggr)^{-n_4} \biggl((P-Q)^2-m_5^2\biggr)^{-n_5} \nonumber \end{eqnarray} and their derivatives $J'$ at $k^2=m_H^2$ or at $k^2=0$. The most difficult is a calculation of the all--massive scalar master integral corresponding to the topology shown in the Fig.~1. \vspace*{0.4cm} \setlength{\unitlength}{1cm} \begin{picture}(15,3) \put(5,0){\epsfig{file=a.eps,height=3cm}} \end{picture} \begin{center} \parbox{6in}{\small\baselineskip=12pt Fig.~1. The two loop all--massive master graph. Solid line represents Higgs bosons. } \end{center} \vspace*{0.4cm} This integral has a discontinuity that is an elliptic integral, resulting from integration over the phase space of three massive particles, and is not expressible in terms of polylogarithms. However one can show \cite{ScharfTausk} that on--shell $k^2=m_H^2$ or at the threshold $k^2=9 m_H^2$ this is not the case. We use the dispersive method \cite{Broadhurst,BaubergerBohm} to evaluate this finite integral on the mass shell: \begin{equation} m_H^2\, J(k^2;m_H^2,m_H^2,m_H^2,m_H^2,m_H^2)= \sigma_a(k^2/m_H^2)+\sigma_b(k^2/m_H^2), \label{master} \end{equation} where $\sigma_{a,b}$ correspond to the dispersive integrals calculated, respectively, from the two-- and three--particle discontinuities, which are itself reduced to one--dimensional integrals. The $\tanh^{-1}$ functions entering $\sigma_{a,b}$ can be removed integrating by parts either in the dispersive integral \cite{Broadhurst}, or in the discontinuity integral \cite{BaubergerBohm}. By interchanging the order of integrations the latter representation gives the three--particle cut contribution $\sigma_b$ as a single integral of logarithmic functions \cite{BaubergerBohm}. After some rather heavy analysis we obtain at $k^2=m_H^2$: \begin{eqnarray} \sigma_a(1)&=&\int_0^1 dy \, \frac{8}{y^{4}-y^{2}+1} \log \left({\frac {\left (y^{2}+1\right )^{2}}{y^{4}+y^{2}+1}}\right) \left [{\frac {\left (y^{4}-1\right )\log (y)}{y}} -{\frac {\pi\,y}{\sqrt {3}}}\right ] \nonumber \\ &=&{\frac {17}{18}}\,\zeta(3)-{\frac {10}{9}}\,\pi \,C+\pi ^{2}\,\log {2} -{\frac {4}{9}}\,\pi ^{2}\,\log {3}, \\ \sigma_b(1)&=&\int_0^1 dy \, 2\, \log \left({\frac {y^{2}+y+1}{y}}\right) \nonumber \\ &&\left [{\frac {\log (y+1)}{y}} +{\frac {{\frac {\pi\,}{\sqrt {3}}}\left (y^{2}-3\,y+1\right ) -2\,\left (y^{2}-1\right ) \log (y)}{y^{4}+y^{2}+1}}-{\frac {\log (y)}{y+1}}\right ] \nonumber\\ &=&{\frac {1}{18}}\zeta(3)+{\frac {4}{9}}\,\pi \,C-\pi ^{2}\,\log {2} +{\frac {4}{9}}\pi ^{2}\,\log {3}. \end{eqnarray} Here $ C = \mbox{Cl}(\pi/3) = \mathop{\mathrm{Im}} \,\mbox{li}_2\left(\exp\left(i\pi/3\right)\right)= 1.01494\: 16064\: 09653\: 62502\dots $ As a result we find \begin{eqnarray} m_H^2\, J(m_H^2;m_H^2,m_H^2,m_H^2,m_H^2,m_H^2)&=& \zeta(3)-{\frac {2}{3}}\,\pi\,C \nonumber \\ &=&-0.92363\: 18265\: 19866\: 53695 \dots \label{HHHHH} \end{eqnarray} The numerical value is in agreement with the one, calculated using the momentum expansion \cite{DavydychevTausk}, and with the numerical values given in \cite{MDR,Adrian}. Given the value (\ref{HHHHH}), the simplest way to calculate the derivative of the integral (\ref{master}) is to use Kotikov's method of differential equations \cite{Kotikov,MDR} \begin{eqnarray} m_H^4\, J'(m_H^2;m_H^2,m_H^2,m_H^2,m_H^2,m_H^2)&=& {\frac {2}{3}}\,\pi\,C-\zeta(3)-{\frac {\pi^{2}}{9}} \nonumber \\ &=&-0.17299\: 08847\: 12284\: 42069 \dots \end{eqnarray} All the other two--loop self energy scalar integrals contain ``light'' particles (Nambu--Goldstone or $W$, $Z$--bosons) and some of them are IR divergent in the limit $M_{W,Z}\to 0$. In principle, one can calculate these IR divergent integrals in Landau gauge, where masses of Nambu--Goldstone bosons are equal to zero and IR divergences are represented as (double) poles at $\epsilon=0$ \cite{MDR}. However, in order to have an additional check of the cancellation in the final answer of all the IR divergent $\log(M_{W,Z}^2)$--terms, we work in 't~Hooft--Feynman gauge. For the infra--red finite integrals the correct answer in the leading order in $m_H^2$ is obtained just by setting $M_{W,Z}=0$. We agree with the results for these integrals given in \cite{MDR,ScharfTausk}. The two--loop IR divergent integrals correspond to the topologies shown in the Fig.~2, which contain ``massless'' propagators squared. \vspace*{0.4cm} \setlength{\unitlength}{1cm} \begin{picture}(15,3) \put(1,0){\epsfig{file=b.eps,height=3cm}} \put(10,0){\epsfig{file=c.eps,height=3cm}} \end{picture} \begin{center} \parbox{6in}{\small\baselineskip=12pt Fig.~2. The two--loop IR divergent graphs. Dashed line represents ``light'' particles.} \end{center} \vspace*{0.4cm} The relevant technique to handle these diagrams follows from the so--called asymptotic operation method \cite{Tkachov}. According to the recipe of As--operation, the formal Taylor expansion in small mass $M_{W}$ entering propagator (or its powers) should be accomplished by adding the terms, containing the $\delta$--function or its derivatives. The additional terms counterbalance the infra--red singularities, arising in the formal expansion of propagator. In our case we have \begin{eqnarray} \frac{1}{(P^{2} - M_{W}^{2})^{2}} & = & \frac{1}{(P^{2})^{2}} + 2\frac{M_{W}^{2}}{(P^{2})^{3}} + ... \nonumber \\ &+& C_{1}(M_{W})\delta^{(d)}(P) + C_{2}(M_{W})\partial^{2}\delta^{(d)}(P)+ ... \end{eqnarray} Here the first coefficient functions $C_{i}(M_{W})$ read \begin{eqnarray} C_{1}(M_{W}) & = & \int D^{(d)}P \frac{1}{(P^{2} - M_{W}^{2})^{2}} \sim {\cal O}(M_{W}^{0}), \\ C_{2}(M_{W}) & = & \frac{1}{2d}\int D^{(d)}P \frac{P^{2}}{(P^{2} - M_{W}^{2})^{2}} \sim {\cal O}(M_{W}^{2}). \nonumber \end{eqnarray} This equality is to be understood in the following sense \cite{Tkachov}. One should integrate both parts of the equation multiplied by a test function and then take the limit $d \rightarrow 4$. If one keeps in the expansion all terms up to order $M_{W}^{2 n}$, the resulting expression will represent a correct expansion of the initial integral to order $o(M_{W}^{2 n})$. To obtain the leading contribution to the diagrams Fig.~2, it suffices just to take the first term of the Taylor expansion and, correspondingly, the first ``counterterm'' $C_{1}\delta^{(d)} (P)$. Finally, the combination of Mellin--Barnes representation and Kotikov's method gives the following answer, corresponding to the first graph in Fig.~2 neglecting the terms of order ${\cal O}(M_W^2/m_H^2)$: \begin{eqnarray} m_H^2\, J(m_H^2;2\,M_W^2,M_W^2,0,M_W^2,m_H^2)&=& {\xi^{2\epsilon}_H}\Biggl ({\frac {2\,i\,\pi}{\epsilon}}-{\frac {i\,\pi}{2}} +\frac{2}{\epsilon}\,\log (\frac{M_W^2}{m_H^2}) \\ &&-\frac{1}{2}+{\frac {5\,\pi^{2}}{6}} -\log (\frac{M_W^2}{m_H^2}) +\frac{1}{2}\log^{2} (\frac{M_W^2}{m_H^2})\Biggr ); \nonumber \\ m_H^4\, J'(m_H^2;2\,M_W^2,M_W^2,0,M_W^2,m_H^2)&=& {\xi^{2\epsilon}_H}\Biggl (-{\frac {2\,i\,\pi}{\epsilon}}+2\,i\,\pi -\frac{2}{\epsilon}\left(1\,+\,\log (\frac{M_W^2}{m_H^2})\right) \\ &&+2-{\frac {5\,\pi^{2}}{6}} +\log (\frac{M_W^2}{m_H^2}) -\frac{1}{2}\log^{2} (\frac{M_W^2}{m_H^2})\Biggr ). \nonumber \end{eqnarray} The integral diverges as $1/\epsilon$, while if we would set $M_W=0$ from the very beginning, it would diverge as $1/\epsilon^2$. The integral corresponding to the second graph in the Fig.~2 up to ${\cal O}(M_W^2/m_H^2)$ is \begin{eqnarray} J(m_H^2;2\,M_W^2,0,0,M_W^2,m_H^2)&=& {\xi^{2\epsilon}_H}\Biggl ({\frac {2}{\epsilon^{2}}} +\frac{1}{\epsilon}\left(2\,\log (\frac{M_W^2}{m_H^2})-1\right) \\ &&+\frac{1}{2}-{\frac {\pi^{2}}{12}} -\log (\frac{M_W^2}{m_H^2}) +\frac{1}{2}\log^{2} (\frac{M_W^2}{m_H^2})\Biggr) \nonumber \end{eqnarray} The two--loop vacuum integrals needed to evaluate the tadpole counterterm ${\delta v^2}^{(2)}$ have been calculated in \cite{vanderBijVeltman}. \section{Results} The analytic results for the two--loop renormalization constants are: \begin{eqnarray} \delta {v^2}^{(2)}&=& \frac{m_H^{4}{\xi^{2\epsilon}_H}}{16\,M_W^2} \left ( {\frac {72}{\epsilon^{2}}} +{\frac {36\,\pi\,\sqrt {3}-84}{\epsilon}} -162-3\,\pi^{2}+60\,\sqrt {3}C \right ); \label{dv2} \\ \frac{{\delta m_H^2}^{(2)}}{m_H^2}&=& \frac{m_H^{4}{\xi^{2\epsilon}_H}}{64\,M_W^4}\Biggl ( {\frac {576}{\epsilon^{2}}} +\frac {144\,\pi\,\sqrt {3}-1014}{\epsilon} \nonumber \\ &&+{\frac {99}{2}}-252\,\zeta(3) +87\,\pi^{2} -219\,\pi\,\sqrt {3} \nonumber \\ &&+156\,\pi\,C +204\,\sqrt {3}C \Biggr); \label{mh2} \\ \delta Z_H^{(2)}&=&\frac{m_H^{4}{\xi^{2\epsilon}_H}}{64\,M_W^4} \Biggl ( {\frac {3}{\epsilon}} - {\frac {75}{4}} - 126\,\zeta(3) \nonumber \\ &&+{\frac {25\,\pi^{2}}{2}}-76\,\pi\,\sqrt {3}+78\,\pi\,C+108\,\sqrt {3}C \Biggr ); \label{zh2} \\ \frac{\delta M_W^{(2)}}{M_W}&=&\delta Z_w^{(2)}= \frac{m_H^{4}{\xi^{2\epsilon}_H}}{64\,M_W^4}\left ( {\frac {3}{\epsilon}}-\frac{3}{8}-{\frac {\pi^{2}}{6}}+{\frac {3\, \pi\,\sqrt {3}}{2}}-6\,\sqrt {3}C \right ). \label{zw2} \end{eqnarray} For comparison we have also calculated these counterterms following the renormalization scheme \cite{GhinculovvanderBij,fermi_G} and keeping the ${\cal O}(\epsilon)$ terms and found complete agreement with their (partly numerical) results. In this scheme the renormalization constants (\ref{dv2}), (\ref{mh2}) look a bit more complicated due to the presence of the additional $\pi \sqrt{3}\log{3}$ terms \begin{eqnarray} \delta {v^2}^{(2)}&=& \frac{m_H^{4}{\xi^{2\epsilon}_H}}{16\,M_W^2} \Biggl ( {\frac {72}{\epsilon^{2}}} +{\frac {36\,\pi\,\sqrt {3}-84}{\epsilon}} +90-12\,\pi^{2}-12\,\sqrt {3}C \\ &&-36\,\pi\,\sqrt{3}+18\,\pi\,\sqrt{3}\,\log{3} \Biggr ); \label{dv2_e}\nonumber \\ \frac{{\delta m_H^2}^{(2)}}{m_H^2}&=& \frac{m_H^{4}{\xi^{2\epsilon}_H}}{64\,M_W^4}\Biggl ( {\frac {576}{\epsilon^{2}}} +\frac {144\,\pi\,\sqrt {3}-1014}{\epsilon} \nonumber \\ &&+{\frac {2439}{2}}-252\,\zeta(3) +63\,\pi^{2} -363\,\pi\,\sqrt {3} \nonumber \\ &&+156\,\pi\,C -84\,\sqrt {3}C+72\,\pi\sqrt{3}\,\log{3} \Biggr). \label{mh2_e} \end{eqnarray} The wave function renormalization constants $Z_{H,w}$ are identical in these two schemes. As an example of the physical quantity, for which all the schemes should give the same result, we consider the two--loop heavy Higgs correction to the fermionic Higgs width \cite{fermi_G,fermi_DKR}. The correction is given by the ratio \begin{eqnarray} \frac{Z_H}{{M_W^2}_0/M_W^2}&=& 1 + 2\frac{g^2}{16\,\pi^2} \left(\delta Z_H^{(1)}-\frac{\delta M_W^{(1)}}{M_W} \right) \nonumber \\ &&+ \frac{g^4}{(16\,\pi^2)^2}\Biggl[ 2\,\frac{\delta M_W^{(1)}}{M_W}\, \left (\frac{\delta M_W^{(1)}}{M_W}-\delta Z_H^{(1)}\right ) +\left (\delta Z_H^{(1)}-\frac{\delta M_W^{(1)}}{M_W}\right )^{2} \nonumber \\ &&+2\,\delta Z_H^{(2)}-2\,\frac{\delta M_W^{(2)}}{M_W} \Biggr]. \end{eqnarray} Substituting (\ref{zh}), (\ref{zh2})--(\ref{zw2}) we find \begin{eqnarray} &&\frac{Z_H}{{M_W^2}_0/M_W^2}= 1\, +\, \frac{1}{8} \frac{g^2}{16\, \pi^2}\frac{m_H^2}{M_W^2}\left( 13 - 2 \pi \sqrt{3}\right) \\ &&+\, \frac{1}{16}\biggl(\frac{g^2}{16\, \pi^2}\frac{m_H^2}{ M_W^2}\biggr)^2 \left(3-63\,\zeta(3)-{\frac {169\,\pi\,\sqrt {3}}{4}} +{\frac {85\,\pi^{2}}{12}}+39\,\pi\,C+57\,\sqrt {3}\,C\right). \nonumber \end{eqnarray} Again, we find complete agreement with the numerical result \cite{fermi_G} and exact agreement with the result \cite{fermi_DKR}, taking into account that their numeric constant $K_5$ is just minus our integral (\ref{HHHHH}). \section*{Acknowledgments} G.J. is grateful to J.J.~van~der~Bij and A.~Ghinculov for valuable discussions. This work was supported in part by the Alexander von Humboldt Foundation and the Russian Foundation for Basic Research grant 96-02-19-464.
proofpile-arXiv_065-492
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} The observed structures of the universe are thought to originate from the gravitational condensation of primordial mass-energy fluctuations whose evolution depends on the power spectrum of the initial fluctuations, the amount and nature of the various matter components present in the early universe and the eventual existence of a cosmological constant. Since most of the acting gravitational masses remain invisible, constraining cosmological scenario from astronomical observations has been one of the most fascinating challenge of the last decades. From the observations of the spatial galaxy distribution (see de Lapparent, this conference), and from dynamical studies of gravitational systems like galaxies or clusters of galaxies, it is possible to infer the amount of dark matter as well as its distribution, or to have insights into the clustering properties of the visible matter in the universe. However, none of these observations measures {\sl directly} the total amount and the distribution of matter: angular or redshift surveys give the distribution of {\sl light} associated with galaxies, and mass estimations of gravitational structures can only be obtained within the assumption that they are simple relaxed or virialised dynamical systems. Dynamics of galaxy velocity fields seems an efficient and promising approach to map directly the potentials responsible from large-scale flows, but catalogs are still poor and some assumptions are still uncertain or poorly understood on a physical basis. Gravitational lensing effects are fortunately a direct probe of deflecting masses in the universe. They allow to determine directly the amount of matter present along the line-of-sight from observed or reconstructed deflection angles. After the discovery of the first multiply imaged quasar (Walsh et al. 1979) and the first observations of gravitational arcs (Soucail et al. 1987, 1988; Lynds \& Petrossian 1986) and arclets (Fort et al. 1988), gravitational lensing rapidly becomes one of the most useful tool for probing dark matter on all scales and cosmological parameter as well. In the following, we summarize what we have learned from strong and weak lensing regimes in clusters of galaxies and how constraints on the cosmological parameters can, or could, be obtained. We will also discuss a technique for measuring gravitational shear that has been developed recently and which opens new perspectives for the observations and the analysis of lensing effect by large-scale structures, as those we discuss in the last section. \begin{figure} \vskip -2cm \hskip 3truecm \psfig{figure=ms2137_modele_color.ps,width=10. cm} \caption{Model of MS2137-23. This cluster is at redshift 0.33 and shows a tangential arc and the first radial arc ever detected. Up to now it is the most constrained cluster and the first one where counter images were predicted before being observed (Mellier et al. 1993). The external and the dark internal solid lines are the critical lines. The internal grey ellipse and the diamond are the caustic lines. The thin isocontours shows the positions of the arcs and their counter images.} \end{figure} \section{Definitions and lensing equations} Let us remind the basic principles of the gravitational lensing effects. A deflecting mass changes the apparent position of a source $\vec \theta_S$ into the apparent image position $\vec \theta_I$ by the quantity $\vec \alpha$: \begin{equation} \vec \theta_S=\vec \theta_I+\vec \alpha(\vec \theta_I) \ . \end{equation} The deflection angle $\vec \alpha(\vec \theta_I)$ is the gradient of the two-dimensional (projected along the line of sight at the angular position $\vec \theta_I$) gravitational potential $\phi$. The gravitational distortion of background objects is described by the Jacobian of the transformation, namely the amplification matrix $\cal A$ between the source and the image plane (Schneider, Ehlers \& Falco 1992): \begin{equation} {\cal A}=\pmatrix{ 1-\kappa-\gamma_1 & -\gamma_2 \cr -\gamma_2 & 1-\kappa+\gamma_1 \cr }, \end{equation} where $\kappa$ is the convergence, $\gamma_1$ and $\gamma_2$ are the shear components. They are related to the Newtonian gravitational potential $\phi$ by: \begin{equation} \kappa={1\over 2} \nabla^2 \phi={\Sigma\over \Sigma_{\rm crit.}}; \ \ \ \gamma_1={1\over 2}(\phi_{,11}-\phi_{,22}) \ ; \ \ \ \gamma_2=\phi_{,12} \ , \end{equation} where $\Sigma$ is the projected mass density and $\Sigma_{\rm crit.}$ is the critical mass density which would exactly focus a light beam originating from the source on the observer plane. It depends on the angular diameter distances $D_{ab}$, (where $a,b=[o(bserver),l(ens) or s(ource)]$) involved in the lens configuration: \begin{equation} \Sigma_{\rm crit.}={c^2\over 4\pi G}{D_{os}\over D_{ol} D_{ls}} \ . \end{equation} Depending on the ratio $l=\Sigma / \Sigma_{\rm crit.}$ we most often distinguish for data analysis the strong lensing ($l\gsim1$) and the weak lensing ($l\ll1$) regimes. Note that the values of the cosmological parameters enter in the relationship between the angular distance and the redshift. This is this dependence that, in some cases, can be used to constrain $\Omega$ or $\lambda$. \section{Arcs and mass in the central region of clusters of galaxies } Arcs and arclets correspond to strong lensing cases with $l\gg1$ and $l \approx 1$ respectively. Giant arcs form at the points where the determinant of the magnification matrix is (close to) infinite. To these points (the critical lines) correspond the caustic lines in the source plane. From the position of the source with respect to the caustic line, one can easily define typical lensing configuration with formation of giant arcs from the merging of two or three images, or radial arcs and ``straight arc'' as well (see Fort \& Mellier 1994, Narayan \& Bartelmann 1996). \begin{table} \hskip 0.3 truecm \psfig{figure=summary_mass_ARC.proc.ps,width=16. cm} \caption{Summary of mass from giant arcs. We only give some typical examples. More details can be found in Fort \& Mellier (1994) and Narayan \& Bartelmann (1996). The scales are expessed in $h_{100}^{-1}$ Kpc.} \end{table} Some of these typical lensing cases have been observed in rich clusters of galaxies. For MS2137-23 (Mellier et al. 1993; see figure 1), A370 (Kneib et al. 1993), Cl0024+1654 (Wallington et al. 1995), A2218 (Kneib et al. 1995), very accurate models with image predictions have been done which provide among the most precise informations we have about the central condensation of dark matter of clusters of galaxies on scale of $\approx 300$$h^{-1}$kpc (table 1). In particular, giant arcs definitely demonstrate that clusters of galaxies are dominated by dark matter, with mass-to-light ratio (M/L) larger than 100, which closely follows the geometry of the diffuse light distribution associated with the brightest cluster members. Furthermore, since arcs occur when $\Sigma /\Sigma_{\rm crit.}\gg1$, clusters of galaxies must be much more concentrated, e.g. with a smaller core radius, than it was expected from their galaxy and X-ray gas distributions. Note that the occurrence of arcs is also enhanced by the existence of additional clumps observed in most of rich clusters which increases the shear substantially (Bartelmann 1995). The direct observations of substructures by using giant arcs confirm that clusters are dynamically young systems, therefore pointing towards a rather high value of $\Omega$. \section{The weak lensing regime and lensing inversion in clusters} \subsection{Lensing inversion and mass reconstruction} Beyond the region of strong lensing, background galaxies are still weakly magnified. The light deviation induces a small increase of their ellipticity in the direction perpendicular to the gradient of the projected potential (shear). Despite the intrinsic ellipticity of the sources and some observational or instrumental effects it is possible to make statistical analysis of this coherent polarization of the images of background galaxies and to recover the mass distribution of the lens. These weakly distorted galaxies are like a background distorted grid which can be used to probe the projected mass density $\Sigma$ of the foreground lens. The shape parameters of the images $M^I$ are related to the shape parameters of the sources $M^S$ by the equation, \begin{figure} \hskip 1truecm \psfig{figure=a1942BMetACF.eps,width=15. cm} \caption{Shear maps around the rich lensing cluster A1942. The CCD images is a 4 hours exposure obtained at the Canada-France-Hawaii Telescope in excellent seeing conditions (0.65"). North is top and east is on the left. On the left panel, the shear is measured by using the Bonnet \& Mellier's method which consists is computing shape parameters of an annular aperture centered of each individual galaxies. On the right panel, the shear is obtained by computing the autocorrelation (ACF) of the images (see section 4.2). Though the shape is almost the same, the signal to noise is higher with the ACF. Note the central pattern which shows the bimodal nature of the mass distribution located on the two brightest cluster members, and the eastern extension of the shear pattern probably due to a substructure.} \end{figure} \begin{equation} M^{S}= {\cal A} M^I {\cal A}, \end{equation} where $M$ are the second order momenta of objects and ${\cal A}$ the magnification matrix. We thus have a relation between the ellipticity of the sources $\vec \epsilon_S$ and the observed ellipticity of the images $\vec \epsilon_I$: \begin{equation} \vec \epsilon_S= {\vec \epsilon_I-\vec g \over 1-\vec g \cdot \vec \epsilon_I} \ ; \ \vec g={\vec \gamma\over 1-\kappa}\ . \end{equation} In particular, if the sources are randomly distributed then their averaged intrinsic ellipticity verifies $\big <\vec \epsilon_S\big >=0$ and we have \begin{equation} \vec g=\big <\vec \epsilon_I\big >\ . \end{equation} Since the main domain of application of the weak lensing is the large-scale distribution of the Dark Matter, at scale above $>0.5$ $h^{-1}$Mpc , the analysis has been mainly focussed on the weak lensing regime, where $(\kappa, \gamma) \ll 1$. The relations between the physical ($\vec\gamma$) and observable ($\vec\epsilon_I$) quantities are then simpler, since we have, \begin{equation} \big <\vec\epsilon_I\big >= \vec\gamma\ . \end{equation} The projected mass density $\Sigma$ of the lens can be obtained from the distortion field by using Eq.(8) and the integration of Eq. (3) (Kaiser \& Squires 1993): \begin{equation} \kappa(\vec \theta_I)={-2\over \pi} \int {\rm d}^2{\vec \theta}\ {\vec\chi(\vec \theta-\vec \theta_I) \over (\vec \theta-\vec \theta_I)^2} \cdot\vec\gamma(\vec \theta_I) +\kappa_0 \ ; \ \ \ \vec \chi(\vec \theta)=({\theta_1^2- \theta_2^2\over \theta^2} \ , \ {2\theta_1 \theta_2 \over \theta^2}), \end{equation} where $\kappa_0$ is the integration constant. In the weak lensing regime, Eqs.(9) provides a mapping of the total projected mass, using the distortion of the background objects. \begin{table} \hskip 2.0 truecm \psfig{figure=summary_mass_WL.proc.ps,width=14. cm} \caption{Summary mass obtained from weak lensing inversion in the literature. The averaged mass-to-light ratio is higher than the one inferred from giant arc (see table 1) which is interpreted as a real increase with distance from the cluster center. Note the strong uncertainty in the case of MS1054 which is due to the strong dependence of mass with the redshift of sources as the redshift of the lens increases. The scales are expressed in $h_{100}^{-1}$ Kpc.} \end{table} Several improvements of the basic theory of lensing inversion have been discussed in details by Seitz \& Schneider (1995, 1996), Schneider (1995) and Kaiser (1996). Note that the objects are also magnified by a factor $\mu$ which, in case of the weak lensing regime, has the form, \begin{equation} \mu=1+2\kappa\ . \end{equation} This gives potentially another independent way to measure the projected mass density $\Sigma$ of a lens using the magnification instead of the distortion. Table 2 gives a summary of the clusters for which lensing inversion have been attempted. When compare with mass distribution inferred from strong lensing, there is a clear trend towards higher $M/L$ when the scale increases. Bonnet et al. (1994) found $M/L$ close to 600 at 2.5 $h^{-1}$Mpc from the cluster center. Note also the remarkable results from Kaiser \& Luppino on a cluster at redshift 0.83, for which the estimated mass strongly depends on the redshifts of the sources. Indeed in general the estimation of the cluster mass using Eq. (9) requires the knowledge of $\Sigma_{\rm crit.}$, which depends on the usually poorly known redshifts of the sources. Though this is not a critical issue for nearby clusters ($z_l<0.2$), because then ${D_{os}/ D_{ls}}\simeq 1$, it could lead to large mass uncertainties for more distant clusters it is the case for MS1054. Actually, even if the redshift of the sources were known, it would still not be possible to get the absolute value of the mass distribution, because possible mass planes of constant density intercepting the line of sight do not change the shear map. Mathematically, this corresponds to the unknown integration constant $\kappa_0$ in Eq.(10). This degeneracy may be broken if one measures the magnification $\mu$ which depends on the mass quantity inside the light beam (Eq.(3)). While the shear measurement does not require any information in the source plane, the magnification measurement needs the observation of a reference (unlensed) field to calibrate the magnification. Broadhurst et al. (1995) proposed to compare the number count $N(m,z)$ and/or $N(m)$ in a lensed and an unlensed field to measure $\mu$. Depending on the value of the slope $S$ of the number count in the reference field, we observe a bias (more objects) or an anti-bias (less objects) in the lensed field. The particular value $S=0.4$ corresponds to the case where the magnification of faint objects is exactly compensated by the dilution of the number count (Eq.(18)). This method was applied on the cluster A1689 (Broadhurst, 1995), but the signal to noise of the detection remains 5 times lower than with the distortion method for a given number of galaxies. The magnification may also be determined by the changes of the image sizes at fixed surface brightness (Bartelmann \& Narayan 1995). The difficulty with these methods is that they required to measure the shape, size and magnitude of very faint objects up to B=28 which depends on the detection threshold, the convolution mask and the local statistical properties of the noise. These remarks led us to propose a new method to analyze the weak lensing effects, based on the auto-correlation function of the pixels in CCD images, which avoids shape parameter measurements of individual galaxies (Van Waerbeke et al. 1996a). It is described in the next subsection. \subsection{The Auto-correlation method} The CCD image is viewed as a density field rather than an image containing delimited objects. The surface brightness, $I(\vec \theta)$, in the image plane in the direction $\vec \theta$ is related to the surface brightness in the source plane $I^{(s)}$ by the relation, \begin{equation} I(\vec \theta)=I^{(s)}({\cal A}\vec \theta), \end{equation} which can be straightforwardly extended to the auto-correlation function (ACF) (e.g. the two-point autocorrelation function of the light distribution in a given area), \begin{equation} \xi(\vec \theta)=\xi^{(s)}({\cal A}\vec \theta)\ . \end{equation} This equation is more meaningful when it is written in the weak lensing regime, \begin{equation} \xi(\vec \theta)=\xi^{(s)}(\theta)-\theta \ \partial_{\theta} \xi^{(s)}(\theta) [1-{\cal A}] \end{equation} since the local ACF, $\xi(\vec \theta)$, now writes as the sum of an isotropic unlensed term, $\xi^{(s)}(\theta)$, an isotropic lens term which depends on $\kappa$, and an anisotropic term which depends on $\gamma_i$. Let us now explore the gravitational lensing information that can be extracted from the shape matrix $\cal M$ of the ACF, \begin{equation} {\cal M}_{ij}={\int {\rm d}^2\theta\ \xi (\vec \theta)\ \theta_i\ \theta_j\over \int {\rm d}^2\theta\ \xi (\vec \theta)}\ . \end{equation} The shape matrix in the image plane is simply related to the shape matrix in the source plane ${\cal M}^{(s)}$ by ${\cal M}_{ij}={\cal A}_{ik}^{-1} {\cal A}_{jl}^{-1} {\cal M}^{(s)}_{kl}$. If the galaxies are isotropically distributed in the source plane, $\xi^{(s)}$ is isotropic, and in that case ${\cal M}^{(s)}_{ij}=M\delta_{ij}$, where $\delta_{ij}$ is the identity matrix. Using the expression of the amplification matrix $\cal A$ we get the general form for $\cal M$, \begin{equation} {\cal M}={M(a+|g|^2)\over (1-\kappa)^2(1-|g|^2)} \pmatrix{ 1+\delta_1 & \delta_2 \cr \delta_2 & 1-\delta_1 \cr }\ . \end{equation} The observable quantities (distortion $\delta_i$ and magnification $\mu$) are given in terms of the components of the shape matrix, \begin{equation} \delta_1={{\cal M}_{11}-{\cal M}_{22}\over {\rm tr}({\cal M})} \ ; \ \ \ \delta_2={2{\cal M}_{12}\over {\rm tr}({\cal M})} \ ; \ \ \ \mu=\sqrt{{\rm det}({\cal M})\over M}, \end{equation} where ${\rm tr}({\cal M})$ is the trace of $\cal M$ and ${\rm det}({\cal M})$ is the determinant of $\cal M$. As for sheared galaxies, we see that the distortion is available from a direct measurement in the image plane while the magnification measurement requires to know the value of $M$ which is related to the light distribution in the source plane, or in an unlensed reference plane. The ACF provides a new and independent way to measure $\delta_i$ and $\mu$ which does not require shape, size or photometry of individual galaxies. Furthermore, the signal to noise ratio is proportional to the number density of background galaxies, $N$, instead as $\sqrt{N}$ for the standard method (see figure 2). A description of its practical implementation and first results are given in Van Waerbeke et al. (1996a) and Van Waerbeke \& Mellier (1996b). Clearly, the ACF is the most powerful technique for measuring the orientation of very weak shear as the one we expect from large-scale structures. \section{The matter distribution on very large scale} The direct observation of the mass distribution on scale larger than 10 $h^{-1}$Mpc (or $\approx $ 1 square degree) is one of the great hopes of the weak gravitational lensing approach. Two observational directions are now being investigated. In the first one we search for the shear of dark condensation of mass that can be responsible for the magnification bias of luminous quasars and radio-sources. It probes the lumpiness of matter distribution within large-scale structures. In the second one, we analyze the statistical properties of weakly lensed background galaxies on degree scales in order to obtain constraints on the cosmological parameters and on the projected power spectrum. \subsection{Shear around radio-sources and the lumpiness of matter} Fugmann (1990) and Bartelmann \& Schneider (1993a,b) have demonstrated that there is a strong correlation between the presence of galaxies or clusters and bright quasars. They interpreted this as a magnification bias induced the galaxy over-densities along the line of sight. However, the measured correlation is too strong to be only due to individual galaxies, so Bartelmann \& Schneider suggested that the magnification bias originates from groups or even rich cluster of galaxies. If this suggestion is correct, the deflector responsible for the magnification bias should also induce gravitational shear onto background sources. \begin{figure} \hskip 3truecm \psfig{figure=q1622imageandshear.ps,width=10. cm} \caption{Shear map around the bright quasar Q1622. The image obtained at CFHT clearly shows a coherent shear. The ellipses shows the center of the shear pattern which is very close to the quasar (dark dot). The others solid lines are galaxy number isodensity contours. Clearly the quasars, the shear pattern and a galaxy concentration are almost positioned at the same place which reinforces the hypothesis of a magnification effect on the quasar.} \end{figure} Bonnet et al. (1993), Mellier et al. (1994) and Van Waerbeke et al. (1996a) found a strong gravitational shear and a galaxy excess around the quasar pair Q2345+007, the shear pattern being associated with a group or a small cluster. In the same way Fort et al. (1996a) have measured the shear around a few over-bright QSO-s that could be magnified by nearby lensing mass (figure 3). It is the first tentative to detect gravitational structures from a mass density criteria rather than luminosity excess. The field of view around each QSO is small and up to now, there is not enough data to draw a synthetic view of the global matter distribution. However, both weak lensing detection around bright radio-sources and the correlations found by Bartelmann \& Schneider seem to favor a model where clumps of dark matter are more numerous that expected and concentrated on groups and clusters of galaxies. \subsection{Statistical analysis of shear on very large scale} In the statistical studies of the shear at very large scale the lenses are not individually identified, but viewed as a random population affecting the shape of the galaxies with an efficiency depending on their distances. Indeed, in this approach, we consider the statistical properties of the shear measured on backgroung objects for randomly chosen line-of-sights. The measured shears are actually filtered at a given angular scale so that a signal of cosmological interest can be extracted. For a filtering scale of about one degree, the structures responsible of the gravitational shear being at a redshift of about 0.4 are expected to be on scales above 10 $h^{-1}$Mpc , that is in a regime where their properties can be easily predicted with the linear or perturbation theory. Within this line of thought, Blandford et al. (1991), Miralda-Escud\'e (1991) and Kaiser (1992) argued that the projected power spectrum should be measurable with such a method provided shape parameters are averaged on the degree scale, as it is illustrated on figure 4. \begin{figure} \hskip 2.0 truecm \psfig{figure=simul.lss.ludo.2.ps,width=14. cm} \caption{ Simulation of shear expected from large scale structures. The left panel shows a 256$^3$ Mpc$^3$ box with dark dots indicating the location of matter. The long filaments are large scale structures which originated from a uniform distribution and an initial power spectrum $P(k)=k^{-1}$ under the adhesion approximation. The right panel shows a slice which represents the projected mass distribution as it would be observed by an observer. The thin straight lines are the local orientation and intensity of the shear. It illustrates what we expect from the future observations of weak lensing with wide field CCD. } \end{figure} For such a scale, however, the shear is expected to be as low as 1\%. Therefore its detection requires high image quality to avoid uncontrolled errors, and large angular coverage so that it is possible to separate the gravitational shear from the intrinsic galactic ellipticities by averaging over thousands of galaxies. So far such a detection has not been completed due to these severe constraints, but there are now projects to build very large CCD mosaic camera which could be capable of doing large deep imaging surveys. The MEGACAM project conducted by the French institutions, CEA (CE in Saclay) and INSU, the Canadian CNRC and the CFHT Corporation consists in building a 16K$\times$16K camera at the prime focus of CFHT. This camera will provide a total field of view of 1$^o\times$1$^o$, with image quality lower than 0.5" on the whole field. Using a Pertubation Theory approach, Bernardeau et al. (1996) have analyzed the statistical properties of the gravitational shear averaged on MEGACAM scales. They have shown that with the observation of about 25 such fields, one could recover the projected power spectrum and $\Omega$ independently by using the variance and the skewness (third moment) of the one-point probability distribution function of the local convergence in the sample \footnote{detailed calculations have shown however that there is a slight degeneracy with the cosmological constant.}. Basically, these moments write, \begin{equation} \big <\kappa^2_{\theta}\big > \propto \ \ P(k)\ \Omega^{1.5} \ \ z_s^{1.5} \end{equation} and \begin{equation} {\big <\kappa^3_{\theta}\big > \over \big <\kappa^2_{\theta}\big >^2} \propto \ \ \Omega^{-0.8} \ \ z_s^{-1.35} \end{equation} where $\theta =30'$ is the scale where the convergence $\kappa$ is averaged, $P(k)$ is the projected power spectrum of the dark matter and $z_s$ is the averaged redshift of sources. Note that the skewness, expressed as the ratio of the third moment by the square of the second, does not depend on $P(k)$ and provides a direct information on $\Omega$. These important results show that a very large-scale survey of gravitational shear can provide important cosmological results which could be compared with those coming from large-scale flows and observations of the cosmic microwave background anisotropy spectrum that are expected to culminate with the COBRAS/SAMBA satellite mission. However, there are two shortcomings that must be handled carefully: first, it requires detection of very weak shear from shape parameters of galaxies. Thus, the image quality of the survey is a crucial issue. The ACF method proposed by Van Waerbeke et al. (1996a) should provide accurate shape information with a high signal to noise ratio. It is probably the best available method for measuring shear on very large scales. The second point is the strong dependence of the variance and the skewness with the redshift of lensed sources. Since there are no hopes to obtain redshifts of these galaxies even from spectroscopy with the VLT-s, we face on a difficult and crucial problem. It may be overcome if multicolor photometric data can provide accurate redshifts for very faint galaxies. Another possible solution could be the analysis of the radial magnification bias of faint distant galaxies around rich clusters which, indeed, probes the redshift distribution of the sources (Fort et al. 1996b and see section 6). \section{Measuring the cosmological constant} One remarkable property of gravitational lensing effect is the local change of the galaxy number density. The observable number density of sources results of the competition between deflection effect which tends to enlarge the projected area and the magnification which increases the number of faint sources. The expected galaxy number density is, \begin{equation} N(r) = N_0\,\mu(r)^{2.5 \alpha -1}, \end{equation} where $\mu$ is the magnification at the position $r$ and $\alpha= {\rm d}\ \log(N) / {\rm d} m$ is the slope of the galaxy counts. When $\alpha$ is larger than 0.4, the galaxy number density increases. At faint limiting magnitude, $\alpha$ becomes significantly smaller than 0.4 and a clear galaxy depletion can be detected (Broadhurst 1995, Fort et al. 1996b). \begin{figure} \hskip .5 truecm \psfig{figure=depletion.blois.full.proc.eps,width=16. cm} \caption{Principle of measurement of redshifts from depletion. The right panel shows the depletions curves expected by a singular isothermal sphere. If the lens is perfectly known the minimum of each depletion curve (right panel) depends only of the redshift of the source and the radial position of the critical line is actually equivalent to a redshift. In a realistic case, the redshift distribution is broad and the left panel shows the depletion as it would be observed: instead of the single peaked depletion we expect a more pronounced minimum between two radii (= two redshifts ) whose angular positions strongly depend on the cosmological constant for high redshift sources. Thus, if the mass distribution of the lens is well known as in the rich lensing cluster Cl0024+1654, $\lambda$ can be inferred from the shape of the depletion curve. } \end{figure} \begin{figure} \psfig{figure=depletion.cl0024.lambda.blois.proc.eps,width=14. cm} \caption{Measuring the cosmological constant from the depletion observed in Cl0024. The left panel shows the depletion curve observed in Cl0024+1654. From the redshift of the first minimum and the second minimum, one can constrain the cosmological constant in order to position the second minimum at its right angular position. Whatever the redshift of the most distant sources visible on the images, we see that the angular position where the depletion curve raises again imposes that $\lambda>0.65$} \end{figure} The shape of the depletion curves depends on the magnification $\mu$ which is a complicated function of the lensing potential, the redshift distribution of sources and the cosmology. For a singular isothermal sphere, the amplification at radius $r$ writes, \begin{equation} \mu(r) = {4 \pi \sigma^2 \over c^2}\,{D_{ls} \over D_{os}}\,{r \over r-1}, \end{equation} and the depletion curve for a single redshift of sources looks like those shown in figure 5. For a redshift distribution, the depletion curve shows a plateau between two extremum points corresponding to the lowest and highest redshift of the sources. Once the redshift of one source is known, the radial position of the highest redshift strongly depends on the cosmological constant. Fort et al. (1996b) tentatively measured depletions curves in Cl0024+1654 (see figure 6) and A370 of faint galaxies close to the noise level. With the hypothesis of a single lens along the line of sight they found that the angular position of the highest-redshift sources is very high and imposes that the most distant galaxies visible in the field have redshifts larger than $2$, while the width of the depleted areas extend as far as 60 arcseconds which is incompatible with a low-$\lambda$ universe. In fact, the observations provide a lower limit $\lambda >0.6$ (figure 6). \section{Conclusion} In the last ten years, the gravitational lensing effects turned out to be among the most promising tools for cosmology. It is indeed a direct probe of the large-scale cosmic mass distribution and some observable quantities revealed to be extremely sensitive to the cosmological parameters. We summarize in table 3 what we learned about $\Omega$ from the mass distribution in rich clusters of galaxies and what we expect in the near future. It shows that we can hope for strong and reliable constraints on $\Omega$ and $\lambda$ from the developing observational tools: we are now aiming at a determination of the cosmological parameters within 10\% accuracy. \begin{table} \hskip 3.0 truecm \psfig{figure=conclu.blois.proc.ps,width=10. cm} \caption{A summary of the values of the cosmological parameters as they are inferred from gravitational lensing. They are still some uncertainties and many hopes for the future. But a real trend toward $\Omega >0.3$ seems well established.} \end{table} \acknowledgements We thanks P. Schneider, for discussions and enthusiastic support. F. Bernardeau is grateful to IAP, where most of this work has been completed, for its hospitality. \section{References} {\parindent=0pt \parskip=3 pt Bartelmann, M., Schneider, P. (1993a) A\&A 259, 413. Bartelmann, M., Schneider, P. (1993b) A\&A 271, 421. Bartelmann, M. (1995) A\&A 299, 661. Bartelmann, M., Narayan, R. (1995) ApJ 451, 60. Bernardeau, F., Van Waerbeke, L., Mellier, Y. (1996) in preparation Blandford, R.D., Saust, A.B., Brainerd, T.G., Villumsen, J.V. (1991), MNRAS, 251, 600 Bonnet, H., Fort, B., Kneib, J-P., Mellier, Y., Soucail, G. (1993) A\&A 280, L7 Bonnet, H., Mellier, Y., Fort, B. (1994) ApJ 427, L83. Bonnet, H., Mellier, Y. (1995) A\&A 303, 331. Broadhurst, T. (1995) SISSA preprint astro-ph/9511150. Broadhurst, T., Taylor, A.N., Peacock, J. (1995) ApJ 438, 49. Fahlman, G., Kaiser, N., Squires, G., Woods, D. (1994) ApJ 437. 56. Fort, B., Prieur, J.-L., Mathez, G., Mellier, Y., Soucail, G. (1988) A\&A 200, L17. Fort, B., Mellier, Y., (1994) A\&A Review 5, 239, 292. Fort, B., Mellier, Y., Dantel-Fort, M., Bonnet, H., Kneib, J.-P. (1996a) A\&A 310, 705. Fort, B., Mellier, Y., Dantel-Fort (1996) SISSA preprint astro-ph/9606039. Fugmann, W. (1990) A\&A 240, 11. Kaiser, N. (1992) ApJ 388, 272. Kaiser, N. (1996) SISSA preprint astro-ph/9509019. Kaiser, N., Squires, G. (1993) ApJ 404, 441 Kneib, J.-P., Mellier, Y., Fort, B., Mathez, G. (1993) A\&A 273, 367. Kneib, J.-P., Mellier, Y., Pell\'o, R., Miralda-Escud\'e, J., Le Borgne, J.-F., Boehringer, H., Picat, J.-P. (1995) A\&A 299, 168. Kassiola, A., Kovner, I., Fort, B. (1992) ApJ 400, 41. Luppino, G., Kaiser, N. (1996) SISSA preprint astro-ph/9601194. Lynds, R., Petrossian, V. (1986) BAAS 18, 1014. Mellier, Y., Fort, B., Kneib, J.-P. (1993) ApJ 407, 33. Mellier, Y., Dantel-Fort, M., Fort, B., Bonnet, H. (1994) A\&A 289, L15. Miralda-Escud\'e, J., (1991) ApJ, 370, 1. Narayan, R., Bartelmann, M. (1996) SISSA preprint astro-ph/9606001 Pierre, M., Le Borgne, J.-F., Soucail, G., Kneib, J.-P. (1996) A\&A 311, 413. Schneider, P., Ehlers, J., Falco, E. E., (1992), {\it Gravitational Lenses}, Springer. Seitz, C., Kneib, J.P., Schneider, P., Seitz, S., (1996) in press. Seitz, C., Schneider, P., (1995) A\&A 302, 9 Seitz, S., Schneider, P., (1996) A\&A 305, 388 Smail, I., Ellis, R.S., Fitchett, M. (1994) MNRAS, 270, 245. Schneider, P. (1995), A\&A, 302, 639 Soucail, G., Fort, B., Mellier, Y., Picat, J.-P. (1987) A\&A 172, L14. Squires, G., Kaiser, N., Babul, A., Fahlmann, G., Woods, D., Neumann, D.M., B\"ohringer, H. (1996a) ApJ 461, 572. Squires, G., Kaiser, N., Falhman, G., Babul, A., Woods, D. (1996b) SISSA preprint astro-ph/9602105. Tyson, J.A., Fisher, P. (1995) ApJL, 349, L1. Van Waerbeke, L., Mellier, Y., Schneider, P., Fort, B., Mathez, G. (1996a) A\&A in press. SISSA preprint astro-ph/9604137. Van Waerbeke, L., Mellier, Y. (1996b). Proceedings of the XXXIst Rencontres de Moriond, Les Arcs, France 1996. SISSA preprint astro-ph/9606100. Wallington, S., Kochanek, C. S., Koo, D. C. (1995) ApJ 441, 58. Walsh, D., Carswell, R.F., Weymann, R.J., (1979), Nature, 279, 381 } \newpage \heading{PROGR\`ES R\'ECENTS EN LENTILLE GRAVITATIONNELLE} \centerline{Y. Mellier$^{(1,4)}$, L. Van Waerbeke$^{(2)}$, F. Bernardeau$^{(3)}$, B. Fort$^{(4)}$} {\it \centerline{$^{(1)}$Institut d'Astrophysique de Paris,} \centerline{98$^{bis}$ Boulevard Arago,} \centerline{75014 Paris, France.} \centerline{$^{(2)}$ Observatoire Midi Pyr\'en\'ees,} \centerline{14 Av. Edouard Belin,} \centerline{31400 Toulouse, France.} \centerline{$^{(3)}$ Service Physique Th\'eorique,} \centerline{CE de Saclay,} \centerline{91191 Gif-sur-Yvette Cedex, France.} \centerline{$^{(4)}$Observatoire de Paris (DEMIRM),} \centerline{61 Av. de l'Observatoire,} \centerline{75014 Paris, France.} } \begin{abstract}{\baselineskip 0.4cm L'effet de lentille gravitationelle est aujourd'hui consid\'er\'e comme un des outils les plus prometteurs de la cosmologie. Il sonde directement la mati\`ere distribu\'ee dans les grandes structures et peut aussi fournir des informations importantes sur les param\`etres cosmologiques, $\Omega$ et $\Lambda$. Dans cet article de revue, nous r\'esumons les progr\'es observationels et th\'eoriques r\'ecents les plus marquants obtenus dans ce domaine au cours des cinq derni\`eres ann\'ees. } \end{abstract} \end{document}
proofpile-arXiv_065-493
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section*{Introduction} Let $Y$ be a connected simplicial complex. Suppose that $\pi$ acts freely and simplicially on $Y$ so that $X=Y/\pi$ is a finite simplicial complex. Let ${\mathcal F}$ a finite subcomplex of $Y$, which is a fundamental domain for the action of $\pi$ on $Y$. We assume that $\pi$ is amenable. The F\o{}lner criterion for amenability of $\pi$ enables one to get, cf. \cite{Ad}, a {\em regular exhaustion} $\big\{Y_{m}\big\}^{\infty}_ {m=1}$, that is a sequence of finite subcomplexes of $Y$ such that (1) $Y_{m}$ consists of $N_{m}$ translates $g.{\mathcal F}$ of ${\mathcal F}$ for $g\in\pi$. (2) $\displaystyle Y=\bigcup^{\infty}_{m=1}Y_{m}\;$. (3) If $\dot{N}_{m,\delta}$ denotes the number of translates of ${\mathcal F}$ which have distance (with respect to the word metric in $\pi$) less than or equal to $\delta$ from a translate of ${\mathcal F}$ having a non-empty intersection with the topological boundary $\partial{Y}_{m}$ of $Y_{m}$ (we identify here $g.{\mathcal F}$ with $g$) then, for every $\delta > 0$, $$ \lim_{m\rightarrow\infty}\;\frac{{\dot{N}}_{m,\delta}}{N_{m}}=0. $$ One of our main results is \begin{theorem}[Amenable Approximation Theorem] $\;$Let $Y$ be a connected simplicial complex. Suppose that $\pi$ is amenable and that $\pi$ acts freely and simplicially on $Y$ so that $X=Y/\pi$ is a finite simplicial complex. Let $\big\{Y_{m}\big\}^{\infty}_ {m=1}$ be a regular exhaustion of $Y$. Then $$ \lim_{m\rightarrow\infty}\;\frac{b^{j}(Y_{m})}{N_{m}}=b_{(2)}^{j}(Y:\pi) \;\;\mbox{ for all }\;\;j\ge 0. $$ $$ \lim_{m\rightarrow\infty}\;\frac{b^{j}(Y_{m}, \partial Y_{m})}{N_{m}}=b_{(2)}^{j}(Y:\pi) \;\;\mbox{ for all }\;\;j\ge 0. $$ \end{theorem} Here $b^{j}(Y_{m})$ denotes the $j^{th}$ Betti number of $Y_m$, $b^{j}(Y_{m}, \partial Y_{m})$ denotes the $j^{th}$ Betti number of $Y_m$ relative its boundary $\partial Y_m$ and $b_{(2)}^{j}(Y:\pi)$ denotes the $j$th $L^2$ Betti number of $Y$. (See the next section for the definition of the $L^2$ Betti numbers of a manifold) \noindent{\bf Remarks.} This theorem proves the main conjecture in the introduction of an earlier paper \cite{DM}. The combinatorial techniques of this paper contrasts with the heat kernel approach used in \cite{DM}. Under the assumption dim $H^{k}(Y)<\infty$, a special case of the Amenable Approximation Theorem above is obtained by combining proofs of Eckmann \cite{Ec} and Cheeger-Gromov \cite{CG}. The assumption dim $H^{k}(Y)<\infty$ is very restrictive and essentially says that $Y$ is a fiber bundle over a $B\pi$ with fiber a space with finite fundamental group. Cheeger-Gromov use this to show that the Euler characteristic of a finite $B\pi$, where $\pi$ contains an infinite amenable normal subgroup, is zero. Eckmann proved the same result in the special case when $\pi$ itself is an infinite amenable group. There is a standing conjecture that any normal covering space of a finite simplicial complex is of determinant class (cf. section 4 for the definition of determinant class and for a more detailed discussion of what follows). Let $M$ be a smooth compact manifold, and $X$ triangulation of $M$. Let $\widetilde M$ be any normal covering space of $M$, and $Y$ be the triangulation of $\widetilde M$ which projects down to $X$. Then on $\widetilde M$, there are two notions of determinant class, one analytic and the other combinatorial. Using results of Efremov \cite{E}, Gromov and Shubin \cite{GS}, one observes as in \cite{BFKM} that the combinatorial and analytic notions of determinant class coincide. It was proved in \cite{BFK}) using estimates of L\"uck \cite{L} that any {\em residually finite} normal covering space of a finite simplicial complex is of determinant class, which gave evidence supporting the conjecture. Our interest in this conjecture stems from work on $L^2$ torsion \cite{CFM}, \cite{BFKM}. The $L^2$ torsion is a well defined element in the determinant line of the reduced $L^2$ cohomology, whenever the covering space is of determinant class. Our next main theorem says that any {\em amenable} normal covering space of a finite simplicial complex is of determinant class, which gives further evidence supporting the conjecture. \begin{theorem}[Determinant Class Theorem] $\;$Let $Y$ be a connected simplicial complex. Suppose that $\pi$ is amenable and that $\pi$ acts freely and simplicially on $Y$ so that $X=Y/\pi$ is a finite simplicial complex. Then $Y$ is of determinant class. \end{theorem} The paper is organized as follows. In the first section, some preliminaries on $L^2$ cohomology and amenable groups are presented. In section 2, the main abstract approximation theorem is proved. We essentially use the combinatorial analogue of the principle of not feeling the boundary (cf. \cite{DM}) in Lemma 2.3 and a finite dimensional result in \cite{L}, to prove this theorem. Section 3 contains the proof of the Amenable Approximation Theorem and some related approximation theorems. In section 4, we prove that an arbitrary {\em amenable} normal covering space of a finite simplicial complex is of determinant class. The second author warmly thanks Shmuel Weinberger for some useful discussions. This paper has been inspired by L\"uck's work \cite{L} on residually finite groups. \section{Preliminaries} Let $\pi$ be a finitely generated discrete group and ${\mathcal U}(\pi)$ be the von Neumann algebra generated by the action of $\pi$ on $\ell^{2}(\pi)$ via the left regular representation. It is the weak (or strong) closure of the complex group algebra of $\pi$, ${\mathbb C}(\pi)$ acting on $\ell^2(\pi)$ by left translation. The left regular representation is then a unitary representation $\rho:\pi\rightarrow{\mathcal U}(\pi)$. Let ${\text{Tr}}_{{\mathcal U}(\pi)}$ be the faithful normal trace on ${\mathcal U}(\pi)$ defined by the inner product ${\text{Tr}}_{{\mathcal U}(\pi)}(A) \equiv (A\delta_e,\delta_e)$ for $A\in{\mathcal U}(\pi)$ and where $\delta_e\in\ell^{2}(\pi)$ is given by $\delta_e(e)=1$ and $\delta_e(g)=0$ for $g\in\pi$ and $g\neq e$. Let $Y$ be a simplicial complex, and $|Y|_j$ denote the set of all $p$-simplices in $Y$. Regarding the orientation of simplices, we use the following convention. For each $p$-simplex $\sigma \in |Y|_j$, we identify $\sigma$ with any other $p$-simplex which is obtained by an {\em even} permutation of the vertices in $\sigma$, whereas we identify $-\sigma$ with any other $p$-simplex which is obtained by an {\em odd} permutation of the vertices in $\sigma$. Suppose that $\pi$ acts freely and simplicially on $Y$ so that $X=Y/\pi$ is a finite simplicial complex. Let ${\mathcal F}$ a finite subcomplex of $Y$, which is a fundamental domain for the action of $\pi$ on $Y$. Consider the Hilbert space of square summable cochains on $Y$, $$ C^j_{(2)}(Y) = \Big\{f\in C^j(Y): \sum_{\sigma\; a\; j-simplex}|f(\sigma)|^2 <\infty \Big\} $$ Since $\pi$ acts freely on $Y$, we see that there is an isomorphism of Hilbert $\ell^2(\pi)$ modules, $$ C^j_{(2)}(Y) \cong C^j(X)\otimes\ell^2(\pi) $$ Here $\pi$ acts trivially on $C^j(X)$ and via the left regular representation on $\ell^2(\pi)$. Let $$ d_{j}:C^{j}_{(2)}(Y)\rightarrow C^{j+1}_{(2)}(Y) $$ denote the coboundary operator. It is clearly a bounded linear operator. Then the (reduced) $L^2$ cohomology groups of $Y$ are defined to be $$ H^j_{(2)}(Y) = \frac{\mbox{ker}(d_j)}{\overline{\mbox{im}(d_{j-1})}}. $$ Let ${d_j}^*$ denote the Hilbert space adjoint of $d_{j}$. One defines the combinatorial Laplacian $\Delta_{j} : C^{j}_{(2)}(Y) \rightarrow C^{j}_{(2)}(Y)$ as $\Delta_j = d_{j-1}(d_{j-1})^{*}+(d_{j})^{*}d_{j}$. By the Hodge decomposition theorem in this context, there is an isomorphism of Hilbert $\ell^2(\pi)$ modules, $$ H^j_{(2)}(Y)\;\; \cong\;\; \mbox{ker} (\Delta_j). $$ Let $P_j: C^{j}_{(2)}(Y)\rightarrow \mbox{ker} \Delta_j$ denote the orthogonal projection to the kernel of the Laplacian. Then the $L^2$ Betti numbers $b_{(2)}^j(Y:\pi)$ are defined as $$ b_{(2)}^j(Y:\pi) = {\text{Tr}}_{{\mathcal U}(\pi)}(P_j). $$ In addition, let $\Delta_j^{(m)}$ denote the Laplacian on the finite dimensional cochain space $C^j(Y_m)$ or $C^j(Y_m,\partial Y_m)$. We do use the same notation for the two Laplacians since all proofs work equally well for either case. Let $D_j(\sigma, \tau) = \left< \Delta_j \delta_\sigma, \delta_\tau\right>$ denote the matrix coefficients of the Laplacian $\Delta_j$ and ${D_j^{(m)}}(\sigma, \tau) = \left< \Delta_j^{(m)} \delta_\sigma, \delta_\tau\right>$ denote the matrix coefficients of the Laplacian $\Delta_j^{(m)}$. Let $d(\sigma,\tau)$ denote the {\em distance} between $\sigma$ and $\tau$ in the natural simplicial metric on $Y$, and $d_m(\sigma,\tau)$ denote the {\em distance} between $\sigma$ and $\tau$ in the natural simplicial metric on $Y_m$. This distance (cf. \cite{Elek}) is defined as follows. Simplexes $\sigma$ and $\tau$ are one step apart, $d(\sigma,\tau)=1$, if they have equal dimensions, $\dim \sigma = \dim \tau =j$, and there exists either a simplex of dimension $j-1$ contained in both $\sigma$ and $\tau$ or a simplex of dimension $j+1$ containing both $\sigma$ and $\tau$. The distance between $\sigma$ and $\tau$ is equal to $k$ if there exists a finite sequence $\sigma = \sigma_0, \sigma_1, \ldots , \sigma_k = \tau$, $d(\sigma_i,\sigma_{i+1})=1$ for $i=0,\ldots,k-1$, and $k$ is the minimal length of such a sequence. Then one has the following, which is an easy generalization of Lemma 2.5 in \cite{Elek} and follows immediately from the definition of combinatorial Laplacians and finiteness of the complex $X=Y/\pi$. \begin{lemma} $D_j(\sigma, \tau) = 0$ whenever $d(\sigma,\tau)>1$ and $ {D_j^{(m)}}(\sigma, \tau) = 0$ whenever $d_m(\sigma,\tau)>1$. There is also a positive constant $C$ independent of $\sigma,\tau$ such that $D_j(\sigma, \tau)\le C$ and ${D_j^{(m)}}(\sigma, \tau)\le C$. \end{lemma} Let $D_j^k(\sigma, \tau) = \left< \Delta_j^k \delta_\sigma, \delta_\tau\right>$ denote the matrix coefficient of the $k$-th power of the Laplacian, $\Delta_j^k$, and $D_j^{(m)k}(\sigma, \tau) = \left< \left(\Delta_j^{(m)}\right)^k \delta_\sigma, \delta_\tau\right>$ denote the matrix coefficient of the $k$-th power of the Laplacian, $\Delta_j^{(m)k}$. Then $$D_j^k(\sigma, \tau) = \sum_{\sigma_1,\ldots\sigma_{k-1} \in |Y|_j} D_j(\sigma, \sigma_1)\ldots D_j(\sigma_{k-1}, \tau)$$ and $$D_j^{(m)k}(\sigma, \tau) = \sum_{\sigma_1,\ldots\sigma_{k-1} \in |Y_m|_j} {D_j^{(m)}}(\sigma, \sigma_1)\ldots {D_j^{(m)}}(\sigma_{k-1}, \tau).$$ Then the following lemma follows easily from Lemma 1.1. \begin{lemma} Let $k$ be a positive integer. Then $D_j^k(\sigma, \tau) = 0$ whenever $d(\sigma,\tau)>k$ and $D_j^{(m)k}(\sigma, \tau) = 0$ whenever $d_m(\sigma,\tau)>k$. There is also a positive constant $C$ independent of $\sigma,\tau$ such that $D_j^k(\sigma, \tau)\le C^k$ and $D_j^{(m)k}(\sigma, \tau)\le C^k$. \end{lemma} Since $\pi$ commutes with the Laplacian $\Delta_j^k$, it follows that \begin{equation}\label{inv} D_j^k(\gamma\sigma, \gamma\tau) = D_j^k(\sigma, \tau) \end{equation} for all $\gamma\in \pi$ and for all $\sigma, \tau \in |Y|_j$. The {\em von Neumann trace} of $\Delta_j^k$ is by definition \begin{equation} \label{vNt} {\text{Tr}}_{{\mathcal U}(\pi)}(\Delta_j^k) = \sum_{\sigma\in |X|_j} D_j^k(\sigma, \sigma), \end{equation} where $\tilde{\sigma}$ denotes an arbitrarily chosen lift of $\sigma$ to $Y$. The trace is well-defined in view of (\ref{inv}). \subsection{Amenable groups} Let $d_1$ be the word metric on $\pi$. Recall the following characterization of amenability due to F\o{}lner, see also \cite{Ad}. \begin{definition} A discrete group $\pi$ is said to be {\em amenable} if there is a sequence of finite subsets $\big\{\Lambda_{k}\big\}^{\infty}_{k=1}$ such that for any fixed $\delta>0$ $$ \lim_{k\rightarrow\infty}\;\frac{\#\{\partial_{\delta}\Lambda_{k}\}}{\# \{\Lambda_{k}\}}=0 $$ where $\partial_{\delta}\Lambda_{k}= \{\gamma\in\pi:d_1(\gamma,\Lambda_{k})<\delta$ and $d_1(\gamma,\pi-\Lambda_{k})<\delta\}$ is a $\delta$-neighborhood of the boundary of $\Lambda_{k}$. Such a sequence $\big\{\Lambda_{k}\big\}^{\infty}_ {k=1}$ is called a {\em regular sequence} in $\pi$. If in addition $\Lambda_{k}\subset\Lambda_{k+1}$ for all $k\geq 1$ and $\displaystyle\bigcup^ {\infty}_{k=1}\Lambda_{k}=\pi$, then the sequence $\big\{\Lambda_{k}\big\}^{\infty}_ {k=1}$ is called a {\em regular exhaustion} in $\pi$. \end{definition} Examples of amenable groups are: \begin{itemize} \item[(1)]Finite groups; \item[(2)] Abelian groups; \item[(3)] nilpotent groups and solvable groups; \item[(4)] groups of subexponential growth; \item[(5)] subgroups, quotient groups and extensions of amenable groups; \item[(6)] the union of an increasing family of amenable groups. \end{itemize} Free groups and fundamental groups of closed negatively curved manifolds are {\em not} amenable. Let $\pi$ be a finitely generated amenable discrete group, and $\big\{\Lambda_{m}\big\}^{\infty}_{m=1}$ a regular exhaustion in $\pi$. Then it defines a regular exhaustion $\big\{Y_m\big\}^{\infty}_{m=1}$ of $Y$. Let $\{P_j(\lambda):\lambda\in[0,\infty)\}$ denote the right continuous family of spectral projections of the Laplacian $\Delta_j$. Since $\Delta_j$ is $\pi$-equivariant, so are $P_j(\lambda) = \chi_{[0,\lambda]}(\Delta_j)$, for $\lambda\in [0,\infty)$. Let $F:[0,\infty)\rightarrow[0,\infty)$ denote the spectral density function, $$ F(\lambda)={\text{Tr}}_{{\mathcal U}(\pi)}(P_j(\lambda)). $$ Observe that the $j$-th $L^{2}$ Betti number of $Y$ is also given by $$ b_{(2)}^j(Y:\pi)=F(0). $$ We have the spectral density function for every dimension $j$ but we do not indicate explicitly this dependence. All our arguments are performed with a fixed value of $j$. Let $E_{m}(\lambda)$ denote the number of eigenvalues $\mu$ of $\Delta_j^{(m)}$ satisfying $\mu\leq\lambda$ and which are counted with multiplicity. We may sometimes omit the subscript $j$ on $\Delta_j^{(m)}$ and $\Delta_j$ to simplify the notation. We next make the following definitions, $$ \begin{array}{lcl} F_{m}(\lambda)& = &\displaystyle\frac{E_{m}(\lambda)}{N_{m}}\;\\[+10pt] \overline{F}(\lambda) & = & \displaystyle\limsup_{m\rightarrow\infty} F_{m}(\lambda) \;\\[+10pt] \mbox{\underline{$F$}}(\lambda) & = & \displaystyle\liminf_{m\rightarrow \infty}F_{m}(\lambda)\; \\[+10pt] \overline{F}^{+}(\lambda) & = & \displaystyle\lim_{\delta\rightarrow +0} \overline{F}(\lambda+\delta) \;\\[+10pt] \mbox{\underline{$F$}}^{+}(\lambda) & = & \displaystyle\lim_{\delta \rightarrow +0}\mbox{\underline{$F$}}(\lambda+\delta). \end{array} $$ \section{Main Technical Theorem} Our main technical result is \begin{theorem} Let $\pi$ be countable, amenable group. In the notation of section 1, one has \begin{itemize} \item[(1)]$\;F(\lambda)=\overline{F}^{+}(\lambda)= \mbox{\underline{$F$}}^{+}(\lambda)$. \item[(2)] $\;\overline{F}$ and {\underline{$F$}} are right continuous at zero and we have the equalities \begin{align*} \overline{F}(0) & =\overline{F}^{+}(0)=F(0)=\mbox{\underline{$F$}} (0)=\mbox{\underline{$F$}}^{+}(0) \\ \displaystyle & =\lim_{m\rightarrow\infty}F_{m}(0)=\lim_{m\rightarrow \infty}\;\frac{\#\{E_{m}(0)\}}{N_{m}}\;. \end{align*} \item[(3)] $\;$Suppose that $0<\lambda<1$. Then there is a constant $K>1$ such that $$ F(\lambda)-F(0)\leq-a\;\frac{\log K^{2}}{\log\lambda}\;. $$ \end{itemize} \end{theorem} To prove this Theorem, we will first prove a number of preliminary lemmas. \begin{lemma} There exists a positive number $K$ such that the operator norms of $\Delta_j$ and of $\Delta_j^{(m)}$ for all $m=1,2\ldots$ are smaller than $K^2$. \end{lemma} \begin{proof} The proof is similar to that in \cite{L}, Lemma 2.5 and uses Lemma 1.1 together with uniform local finiteness of $Y$. More precisely we use the fact that the number of $j$-simplexes in $Y$ at distance at most one from a $j$-simplex $\sigma$ can be bounded \emph{independently} of $\sigma$, say $\#\{\tau \in |Y|_j : d(\tau, \sigma) \leq 1\} \leq b$. \emph {A fortiori} the same is true (with the same constant $a$) for $Y_m$ for all $m$. We now estimate the $\ell^2$ norm of $\Delta \kappa$ for a cochain $\kappa = \sum_\sigma a_\sigma \sigma$ (having identified a simplex $\sigma$ with the dual cochain). Now $$ \Delta \kappa = \sum_\sigma \left ( \sum_\tau D(\sigma, \tau ) a_\tau \right ) \sigma $$ so that $$ \sum_\sigma \left ( \sum_\tau D(\sigma, \tau) a_\tau \right )^2 \leq \sum_\sigma \left ( \sum_{d(\sigma, \tau )\leq 1} D(\sigma, \tau )^2 \right ) \left ( \sum_{d(\sigma, \tau )\leq 1} a_\tau^2 \right ) \leq C^2 b \sum_\sigma \sum_{d(\sigma, \tau ) \leq 1} a_\tau^2, $$ where we have used Lemma 1.1 and Cauchy-Schwartz inequality. In the last sum above, for every simplex $\sigma$, $a_{\sigma}^2$ appears at most $b$ times. This proves that $\|\Delta \kappa \|^2 \leq C^2 b^2 \|\kappa \|^2$. Identical estimate holds (with the same proof) for $\Delta^{(m)}$ which yields the lemma if we set $K=\sqrt{C b}$. \end{proof} Observe that $\Delta_j$ can be regarded as a matrix with entries in ${\mathbb Z} [\pi]$, since by definition, the coboundary operator $d_j$ is a matrix with entries in ${\mathbb Z} [\pi]$, and so is its adjoint $d_j^*$ as it is equal to the simplicial boundary operator. There is a natural trace for matrices with entries in ${\mathbb Z} [\pi]$, viz. $$ {\text{Tr}}_{{\mathbb Z} [\pi]}(A)= \sum_i {\text{Tr}}_{{\mathcal U} [\pi]}(A_{i,i}). $$ \begin{lemma} $\;$Let $\pi$ be an amenable group and let $p(\lambda) = \sum_{r=0}^d a_r \lambda^r$ be a polynomial. Then, $$ {\text{Tr}}_{{\mathbb Z} [\pi]}(p(\Delta_j))= \lim_{m\rightarrow\infty}\frac{1}{N_{m}}\; {\text{Tr}}_{{\mathbb C}}\Big(p\Big(\Delta_j^{(m)}\Big)\Big). $$ \end{lemma} \begin{proof} First observe that if $\sigma\in |Y_m|_j$ is such that $d(\sigma , \partial Y_m) > k$, then Lemma 1.2 implies that $$ D_j^k(\sigma, \sigma) = \left<\Delta_j^k \delta_\sigma, \delta_\sigma\right> = \left<\Delta_j^{(m)k} \delta_\sigma, \delta_\sigma\right> = D_j^{(m)k}(\sigma, \sigma). $$ By (\ref{inv}) and (\ref{vNt}) $$ {\text{Tr}}_{{\mathbb Z} [\pi]}(p(\Delta_j))= \frac{1}{N_m} \sum_{\sigma\in |Y_m|_j} \left< p(\Delta_j)\sigma , \sigma \right>. $$ Therefore we see that $$ \left| {\text{Tr}}_{{\mathbb Z} [\pi]}(p(\Delta_j)) - \frac{1}{N_{m}}\; {\text{Tr}}_{{\mathbb C}}\Big(p\Big(\Delta_j^{(m)}\Big)\Big)\right| \le $$ $$ \frac{1}{N_{m}} \, \sum_{r=0}^d \, |a_r| \sum_{ \begin{array}{lcl} & \sigma\in |Y_m|_j \\ & d(\sigma, \partial Y_m) \leq d \end{array}} \, \left( D^r(\sigma, \sigma) + D^{(m)r}(\sigma, \sigma)\right). $$ Using Lemma 1.2, we see that there is a positive constant $C$ such that $$ \left| {\text{Tr}}_{{\mathbb Z} [\pi]}(p(\Delta_j)) - \frac{1}{N_{m}}\; {\text{Tr}}_{{\mathbb C}}\Big(p\Big(\Delta_j^{(m)}\Big)\Big)\right| \le 2\, \frac{\dot{N}_{m,d}}{N_{m}} \, \sum_{r=0}^d \, |a_r| \, C^r. $$ The proof of the lemma is completed by taking the limit as $m\rightarrow\infty$. \end{proof} We next recall the following abstract lemmata of L\"uck \cite{L}. \begin{lemma} $\;$Let $p_{n}(\mu)$ be a sequence of polynomials such that for the characteristic function of the interval $[0,\lambda]$, $\chi_{[0,\lambda]} (\mu)$, and an appropriate real number $L$, $$ \lim_{n\rightarrow\infty}p_{n}(\mu)=\chi_{[0,\lambda]}(\mu)\;\;\mbox{ and } \;\;|p_{n}(\mu)|\leq L $$ holds for each $\mu\in[0,||\Delta_j||^{2}]$. Then $$ \lim_{n\rightarrow\infty}{\text{Tr}}_{{\mathbb Z} [\pi]}(p_{n}(\Delta_j))=F(\lambda). $$\end{lemma} \begin{lemma} $\;$Let $G:V\rightarrow W$ be a linear map of finite dimensional Hilbert spaces $V$ and $W$. Let $p(t)=\det(t-G^{*}G)$ be the characteristic polynomial of $G^{*}G$. Then $p(t)$ can be written as $p(t)=t^{k}q(t)$ where $q(t)$ is a polynomial with $q(0)\neq 0$. Let $K$ be a real number, $K\geq\max\{1,||G||\}$ and $C>0$ be a positive constant with $|q(0)|\geq C>0$. Let $E(\lambda)$ be the number of eigenvalues $\mu$ of $G^{*}G$, counted with multiplicity, satisfying $\mu\leq\lambda$. Then for $0<\lambda<1$, the following estimate is satisfied. $$ \frac{ \{E(\lambda)\}- \{E(0)\}}{\dim_{{\mathbb C}}V}\leq \frac{-\log C}{\dim_{{\mathbb C}}V(-\log\lambda)}+ \frac{\log K^{2}}{-\log\lambda}\;. $$\end{lemma} \begin{proof}[Proof of theorem 2.1] Fix $\lambda\geq 0$ and define for $n\geq 1$ a continuous function $f_{n}:{\mathbb R}\rightarrow {\mathbb R}$ by $$ f_{n}(\mu)=\left\{\begin{array}{lcl} 1+\frac{1}{n} & \mbox{ if } & \mu\leq\lambda\\[+7pt] 1+\frac{1}{n}-n(\mu-\lambda) & \mbox{ if } & \lambda\leq\mu\leq\lambda+\frac{1}{n} \\[+7pt] \frac{1}{n} & \mbox{ if } & \lambda+\frac{1}{n}\leq \mu \end{array}\right. $$ Then clearly $\chi_{[0,\lambda]}(\mu)<f_{n+1}(\mu)<f_{n}(\mu)$ and $f_{n} (\mu)\rightarrow\chi_{[0,\lambda]}(\mu)$ as $n\rightarrow\infty$ for all $\mu\in[0,\infty)$. For each $n$, choose a polynomial $p_{n}$ such that $\chi_{[0,\lambda]}(\mu)<p_{n}(\mu)<f_{n}(\mu)$ holds for all $\mu\in[0,K^{2}]$. We can always find such a polynomial by a sufficiently close approximation of $f_{n+1}$. Hence $$ \chi_{[0,\lambda]}(\mu)<p_{n}(\mu)<2 $$ and $$ \lim_{n\rightarrow\infty}p_{n}(\mu)=\chi_{[0,\lambda]}(\mu) $$ for all $\mu\in [0,K^{2}]$. Recall that $E_{m}(\lambda)$ denotes the number of eigenvalues $\mu$ of $\Delta_j^{(m)}$ satisfying $\mu\leq\lambda$ and counted with multiplicity. Note that $||\Delta_j^{(m)} || \leq K^{2}$ by Lemma 2.2. $$ \begin{array}{lcl} \displaystyle\frac{1}{N_{m}}\; {\text{Tr}}_{{\mathbb C}}\big(p_{n}(\Delta_j^{(m)})\big) &=&\displaystyle \frac{1}{N_{m}}\sum_{\mu\in [0,K^2]}p_{n}(\mu)\\[+12pt] \displaystyle & =& \displaystyle\frac{ E_{m}(\lambda)}{N_{m}}+\frac{1}{N_{m}}\left\{ \sum_{\mu\in [0,\lambda ]}(p_{n}(\mu)-1)+\sum_{\mu\in (\lambda , \lambda + 1/n]}p_{n}(\mu)\right.\\[+12pt] \displaystyle& &\displaystyle \hspace*{.5in}\left.+\;\sum_{\mu\in (\lambda + 1/n, K^2]}p_{n}(\mu)\right\} \end{array} $$ Hence, we see that \begin{equation} \label{A} F_{m}(\lambda)=\frac{ E_{m}(\lambda)}{N_{m}} \leq\frac{1}{N_{m}}\;{\text{Tr}}_{{\mathbb C}}\big(p_{n}(\Delta_j^{(m)})\big). \end{equation} In addition, $$ \begin{array}{lcl} \displaystyle\frac{1}{N_{m}}\; {\text{Tr}}_{{\mathbb C}}\big(p_{n}(\Delta_j^{(m)})\big)& \leq & \displaystyle\frac{ E_{m}(\lambda)}{N_{m}} +\;\frac{1}{N_{m}}\sup\{p_{n}(\mu)-1:\mu\in[0,\lambda]\}\; E_{m}(\lambda) \\[+16pt] \displaystyle &+&\displaystyle\;\frac{1}{N_{m}}\sup\{p_{n}(\mu): \mu\in[\lambda,\lambda+1/n]\}\; (E_{m}(\lambda+1/n)-E_{m}(\lambda)) \\[+16pt] \displaystyle &+&\displaystyle\;\frac{1}{N_{m}}\sup\{p_{n}(\mu): \mu\in[\lambda+1/n,\;K^{2}]\}\; (E_{m}(K^{2})-E_{m}(\lambda+1/n)) \\[+16pt] \displaystyle &\leq &\displaystyle\frac{ E_{m}(\lambda)}{N_{m}}+ \frac{ E_{m}(\lambda)}{nN_{m}}+ \frac{(1+1/n) (E_{m}(\lambda+1/n)-E_{m}(\lambda))}{N_{m}} \\[+16pt] \displaystyle & &\displaystyle\hspace*{.5in}+\; \frac{(E_{m}(K^{2})-E_{m}(\lambda+1/n))} {nN_{m}} \\[+16pt] \displaystyle &\leq &\displaystyle \frac{ E_{m}(\lambda+1/n)}{N_{m}}+\frac{1}{n}\; \frac{ E_{m}(K^{2})}{N_{m}} \\[+16pt] \displaystyle &\leq & \displaystyle F_{m}(\lambda+1/n)+\frac{a}{n} \end{array} $$ since $E_m(K^2)=\dim C^j(Y_m) \leq aN_m$ for a positive constant $a$ independent of $m$. It follows that \begin{equation} \label{B} \frac{1}{N_{m}}\;{\text{Tr}}_{{\mathbb C}}\big(p_{n}(\Delta_j^{(m)})\big)\leq F_{m} (\lambda+1/n)+\frac{a}{n}. \end{equation} Taking the limit inferior in (\ref{B}) and the limit superior in (\ref{A}), as $m\rightarrow\infty$, we get that \begin{equation} \label{C} {\overline{F}}(\lambda)\leq {\text{Tr}}_{{\mathbb Z} [\pi]}\big(p_{n}(\Delta_j)\big) \leq\mbox{\underline{$F$}}(\lambda+1/n)+\frac{a}{n}. \end{equation} Taking the limit as $n\rightarrow\infty$ in (\ref{C}) and using Theorem 2.4, we see that $$ {\overline{F}}(\lambda)\leq F(\lambda) \leq\mbox{\underline{$F$}}^{+}(\lambda). $$ For all $\varepsilon>0$ we have $$ F(\lambda)\leq\mbox{\underline{$F$}}^{+}(\lambda)\leq\mbox{\underline{$F$}} (\lambda+\varepsilon)\leq {\overline{F}}(\lambda+\varepsilon) \leq F(\lambda+\varepsilon). $$ Since $F$ is right continuous, we see that $$ F(\lambda)={\overline{F}}^{+}(\lambda)=\mbox{\underline{$F$}}^{+} (\lambda) $$ proving the first part of theorem 2.1. Next we apply theorem 2.5 to $\Delta_j^{(m)}$. Let $p_{m}(t)$ denote the characteristic polynomial of $\Delta_j^{(m)}$ and $p_{m}(t)=t^{r_{m}}q_{m}(t)$ where $q_{m}(0)\neq 0$. The matrix describing $\Delta_j^{(m)}$ has integer entries. Hence $p_{m}$ is a polynomial with integer coefficients and $|q_{m}(0)|\geq 1$. By Lemma 2.2 and Theorem 2.5 there are constants $K$ and $C=1$ independent of $m$, such that $$ \frac{F_{m}(\lambda)-F_{m}(0)}{a}\leq\frac{\log K^{2}} {-\log\lambda} $$ That is, \begin{equation}\label{D} F_{m}(\lambda)\leq F_{m}(0)-\frac{a\log K^{2}}{\log\lambda}. \end{equation} Taking limit inferior in (\ref{D}) as $m\rightarrow\infty$ yields $$ \mbox{\underline{$F$}}(\lambda)\leq\mbox{\underline{$F$}}(0) -\frac{a\log K^{2}}{\log\lambda}. $$ Passing to the limit as $\lambda\rightarrow +0$, we get that $$ \mbox{\underline{$F$}}(0)=\mbox{\underline{$F$}}^{+}(0) \qquad \qquad \mbox{and } \qquad\qquad {\overline{F}}(0)={\overline{F}}^{+}(0). $$ We have seen already that ${\overline{F}}^{+}(0)=F(0)=\mbox{\underline{$F$}} (0)$, which proves part ii) of Theorem 2.1. Since $\displaystyle- \frac{a\log K^{2}}{\log\lambda}$ is right continuous in $\lambda$, $$ {\overline{F}}^{+}(\lambda)\leq F(0)-\frac{a\log K^{2}}{\log\lambda}. $$ Hence part iii) of Theorem 2.1 is also proved. \end{proof} We will need the following lemma in the proof of Theorem 0.2 in the last section. We follow the proof of Lemma 3.3.1 in \cite{L}. \begin{lemma} $$ \int_{0+}^{K^2}\left\{\frac{F(\lambda)-F(0)}{\lambda}\right\} d\lambda\le \liminf_{m\to\infty} \int_{0+}^{K^2}\left\{\frac{F_m(\lambda)-F_m(0)}{\lambda}\right\} d\lambda $$ \end{lemma} \begin{proof} By Theorem 2.1, and the monotone convergence theorem, one has \begin{align*} \int_{0+}^{K^2}\left\{\frac{F(\lambda)-F(0)}{\lambda}\right\}d\lambda & = \int_{0+}^{K^2}\left\{\frac{\underline{F}(\lambda)- \underline{F}(0)}{\lambda}\right\}d\lambda\\ & = \int_{0+}^{K^2}\liminf_{m\to\infty}\left\{\frac{F_m(\lambda)- F_m(0)}{\lambda}\right\}d\lambda \\ & = \int_{0+}^{K^2}\lim_{m\to\infty}\left(\inf\left\{\frac{F_n(\lambda)- F_n(0)}{\lambda}|n \ge m\right\}\right)d\lambda \\ & = \lim_{m\to\infty}\int_{0+}^{K^2}\inf\left\{\frac{F_n(\lambda)- F_n(0)}{\lambda}|n \ge m\right\}d\lambda \\ & \le \liminf_{m\to\infty} \int_{0+}^{K^2}\left\{\frac{F_m(\lambda)- F_m(0)}{\lambda}\right\}d\lambda. \end{align*} \end{proof} \section{Proofs of the main theorems} In this section, we will prove the Amenable Approximation Theorem (Theorem 0.1) of the introduction. We will also prove some related spectral results. \begin{proof}[Proof of Theorem 0.1 (Amenable Approximation Theorem)] Observe that \begin{align*} \frac{b^j(Y_{m})}{N_{m}} & = \frac{\dim_{\mathbb C}\Big(\ker(\Delta_j^{(m)})\Big)}{N_{m}} \\ &=F_{m}(0). \end{align*} Also observe that \begin{align*} b_{(2)}^j(Y:\pi) & = \dim_{\pi}\Big(\ker(\Delta_j)\Big)\\ &=F(0). \end{align*} Therefore Theorem $0.1$ follows from Theorem 2.1 after taking the limit as $m\to\infty$. \end{proof} Suppose that $M$ is a compact Riemannian manifold and $\Omega^{j}_{(2)} (\widetilde{M})$ denote the Hilbert space of square integrable $j$-forms on a normal covering space $\widetilde{M}$, with transformation group $\pi$. The Laplacian ${\widetilde{\Delta}}_{j}:\Omega^{j}_{(2)} (\widetilde{M})\rightarrow\Omega^{j}_{(2)}(\widetilde{M})$ is essentially self-adjoint and has a spectral decomposition $\{{\widetilde P}_{j}(\lambda):\lambda\in [0,\infty)\}$ where each ${\widetilde P}_{j}(\lambda)$ has finite von Neumann trace. The associated von Neumann spectral density function, ${\widetilde F}(\lambda)$ is defined as $$ {\widetilde F}:[0,\infty)\rightarrow [0,\infty),\;\;\; {\widetilde F}(\lambda)= {\text{Tr}}_{{\mathcal U}(\pi)} ({\widetilde P}_{j}(\lambda)). $$ Note that ${\widetilde F}(0)=b_{(2)}^{j}(\widetilde{M}:\pi)$ and that the spectrum of $\widetilde{\Delta}_{j}$ has a gap at zero if and only if there is a $\lambda>0$ such that $$ {\widetilde F}(\lambda)={\widetilde F}(0). $$ Suppose that $\pi$ is an amenable group. Fix a triangulation $X$ on $M$. Then the normal cover $\widetilde{M}$ has an induced triangulation $Y$. Let $Y_{m}$ denote be a subcomplex of $Y$ such that $\big\{Y_{m}\big\}^{\infty}_{m=1}$ is a regular exhaustion of $Y$. Let $\Delta^{(m)}_j:C^{j}(Y_{m}, {\mathbb C})\rightarrow C^{j}(Y_{m},{\mathbb C})$ denote the combinatorial Laplacian, and let $E^{(m)}_j(\lambda)$ denote the number of eigenvalues $\mu$ of $\Delta^{(m)}_j$ which are less than or equal to $\lambda$. Under the hypotheses above we prove the following. \begin{theorem}[Gap criterion] $\;$The spectrum of ${\widetilde{\Delta}}_j$ has a gap at zero if and only if there is a $\lambda>0$ such that $$ \lim_{m\rightarrow\infty}\; \frac{E^{(m)}_j(\lambda)-E^{(m)}_j(0)}{N_{m}}=0. $$\end{theorem} \begin{proof} Let $\Delta_j:C^{j}_{(2)}(Y)\rightarrow C^{j}_{(2)}(Y)$ denote the combinatorial Laplacian acting on $L^2$ j-cochains on $Y$. Then by \cite{GS}, \cite{E}, the von Neumann spectral density function $F$ of the combinatorial Laplacian $\Delta_j$ and the von Neumann spectral density function $\widetilde F$ of the analytic Laplacian ${\widetilde{\Delta}}_j$ are dilatationally equivalent, that is, there are constants $C>0$ and $\varepsilon>0$ independent of $\lambda$ such that for all $\lambda\in(0,\varepsilon)$, \begin{equation} \label{star} F(C^{-1}\lambda)\leq {\widetilde F}(\lambda)\leq F(C\lambda). \end{equation} Observe that $\frac{E^{(m)}_j(\lambda)}{N_{m}} = F_m(\lambda)$. Therefore the theorem also follows from Theorem 2.1. \end{proof} There is a standing conjecture that the Novikov-Shubin invariants of a closed manifold are positive (see \cite{E}, \cite{ES} and \cite{GS} for its definition). The next theorem gives evidence supporting this conjecture, at least in the case of amenable fundamental groups. \begin{theorem}[Spectral density estimate] $\;$There are constants $C>0$ and $\varepsilon>0$ independent of $\lambda$, such that for all $\lambda\in (0,\varepsilon)$ $$ {\widetilde F}(\lambda)-{\widetilde F}(0)\leq\frac{C}{-\log(\lambda)}\;. $$\end{theorem} \begin{proof} This follows from Theorem 2.1 and Theorem 3.1 since $\widetilde{\Delta}_j$ has a gap at zero if and only if $ {\widetilde F}_j(\lambda)={\widetilde F}_j(0) $ for some $\lambda>0$. \end{proof} \section{On the determinant class conjecture} There is a standing conjecture that any normal covering space of a finite simplicial complex is of determinant class. Our interest in this conjecture stems from our work on $L^2$ torsion \cite{CFM}, \cite{BFKM}. The $L^2$ torsion is a well defined element in the determinant line of the reduced $L^2$ cohomology, whenever the covering space is of determinant class. In this section, we use the results of section 2 to prove that any {\em amenable} normal covering space of a finite simplicial complex is of determinant class. Recall that a covering space $Y$ of a finite simplicial complex $X$ is said to be of {\em determinant class} if, for $0 \le j \le n,$ $$ - \infty < \int^1_{0^+} \log \lambda d F (\lambda),$$ where $F(\lambda)$ denotes the von Neumann spectral density function of the combinatorial Laplacian $\Delta_j$ as in Section 2. Suppose that $M$ is a compact Riemannian manifold and $\Omega^{j}_{(2)} (\widetilde{M})$ denote the Hilbert space of square integrable $j$-forms on a normal covering space $\widetilde{M}$, with transformation group $\pi$. The Laplacian ${\widetilde{\Delta}}_{j}:\Omega^{j}_{(2)} (\widetilde{M})\rightarrow\Omega^{j}_{(2)}(\widetilde{M})$ is essentially self-adjoint and the associated von Neumann spectral density function, ${\widetilde F}(\lambda)$ is defined as in section 3. Note that ${\widetilde F}(0)=b_{(2)}^{j}(\widetilde{M}:\pi)$ Then $\widetilde{M}$ is said to be of {\em analytic-determinant class}, if, for $0 \le j \le n,$ $$ - \infty < \int^1_{0^+} \log \lambda d {\widetilde F} (\lambda),$$ where ${\widetilde F}(\lambda)$ denotes the von Neumann spectral density function of the analytic Laplacian ${\widetilde{\Delta}}_{j}$ as above. By results of Gromov and Shubin \cite{GS}, the condition that $\widetilde{M}$ is of analytic-determinant class is independent of the choice of Riemannian metric on $M$. Fix a triangulation $X$ on $M$. Then the normal cover $\widetilde{M}$ has an induced triangulation $Y$. Then $\widetilde{M}$ is said to be of {\em combinatorial-determinant class} if $Y$ is of determinant class. Using results of Efremov \cite{E}, and \cite{GS} one sees that the condition that $\widetilde{M}$ is of combinatorial-determinant class is independent of the choice of triangulation on $M$. Using again results of \cite{E} and \cite{GS}, one observes as in \cite{BFKM} that the combinatorial and analytic notions of determinant class coincide, that is $\widetilde{M}$ is of combinatorial-determinant class if and only if $\widetilde{M}$ is of analytic-determinant class. The appendix of \cite{BFK} contains a proof that every residually finite covering of a compact manifold is of determinant class. Their proof is based on L\"uck's approximation of von Neumann spectral density functions \cite{L}. Since an analogous approximation holds in our setting (cf. Section 2), we can apply the argument of \cite{BFK} to prove Theorem 0.2. \begin{proof}[Proof of Theorem 0.2 (Determinant Class Theorem)] Recall that the \emph{normalized} spectral density functions $$ F_{m} (\lambda) = \frac 1 {N_m} E_j^{(m)} (\lambda) $$ are right continuous. Observe that $F_{m}(\lambda)$ are step functions and denote by ${\det}' \Delta_j^{(m)}$ the modified determinant of $\Delta_j^{(m)}$, i.e. the product of all {\em nonzero} eigenvalues of $\Delta_j^{(m)}$. Let $a_{m, j}$ be the smallest nonzero eigenvalue and $b_{m, j}$ the largest eigenvalue of $\Delta_j^{(m)}$. Then, for any $a$ and $b$, such that $0 < a < a_{m, j}$ and $b > b_{m,j}$, \begin{equation}\label{one} \frac 1 {N_m} \log {\det}' \Delta_j^{(m)} = \int_a^b \log \lambda d F_{m} (\lambda). \end{equation} Integration by parts transforms the Stieltjes integral $\int_a^b \log \lambda d F_{m} (\lambda)$ as follows. \begin{equation}\label{two} \int_a^b \log \lambda d F_{m} (\lambda) = (\log b) \big( F_{m} (b) - F_{m} (0) \big) - \int_a^b \frac {F_{m} (\lambda) - F_{m} (0)} \lambda d \lambda. \end{equation} As before, $F(\lambda)$ denotes the spectral density function of the operator $\Delta_j$ for a fixed $j$. Recall that $F(\lambda)$ is continuous to the right in $\lambda$. Denote by ${\det}'_\pi\Delta_j$ the modified Fuglede-Kadison determinant (cf. \cite{FK}) of $\Delta_j$, that is, the Fuglede-Kadison determinant of $\Delta_j$ restricted to the orthogonal complement of its kernel. It is given by the following Lebesgue-Stieltjes integral, $$ \log {\det}^\prime_\pi \Delta_j = \int^{K^2}_{0^+} \log \lambda d F (\lambda) $$ with $K$ as in Lemma 2.2, i.e. $ || \Delta_j || < K^2$, where $||\Delta_j||$ is the operator norm of $\Delta_j$. Integrating by parts, one obtains \begin{align} \label{three} \log {\det}^\prime_\pi (\Delta_j) & = \log K^2 \big( F(K^2) - F(0) \big) \nonumber\\ & + \lim_{\epsilon \rightarrow 0^+} \Big\{(- \log \epsilon) \big( F (\epsilon) -F(0) \big) - \int_\epsilon^{K^2} \frac {F (\lambda) - F (0)} \lambda d \lambda \Big\}. \end{align} Using the fact that $ \liminf_{\epsilon \rightarrow 0^+} (- \log \epsilon) \big( F (\epsilon) - F (0) \big) \ge 0$ (in fact, this limit exists and is zero) and $\frac {F (\lambda) - F (0)} \lambda \ge 0$ for $\lambda > 0,$ one sees that \begin{equation}\label{four} \log {\det}^\prime_\pi (\Delta_j) \ge ( \log K^2) \big(F (K^2) - F(0) \big) - \int_{0^+}^{K^2} \frac {F(\lambda) - F (0)} \lambda d \lambda. \end{equation} We now complete the proof of Theorem 0.2. The main ingredient is the estimate of $\log {\det}'_\pi(\Delta_j)$ in terms of $\log {{\det}}^\prime \Delta_j^{(m)}$ combined with the fact that $\log {\det}' \Delta_j^{(m)}\ge 0$ as the determinant $\det' \Delta_j^{(m)}$ is a positive integer. By Lemma 2.2, there exists a positive number $K$, $1 \le K < \infty$, such that, for $m \ge 1$, $$ || \Delta_j^{(m)} || \le K^2 \quad {\text{and}}\quad || \Delta_j || \le K^2.$$ By Lemma 2.6, \begin{equation}\label{five} \int_{0^+}^{K^2} \frac {F (\lambda) - F (0)} \lambda d \lambda \le \liminf_{m \rightarrow \infty} \int_{0^+}^{K^2} \frac {F_{m} (\lambda) - F_{m} (0)} \lambda d \lambda. \end{equation} Combining (\ref{one}) and (\ref{two}) with the inequalities $\log {\det}' \Delta_j^{(m)} \ge 0$, we obtain \begin{equation}\label{six} \int_{0^+}^{K^2} \frac {F_{m} (\lambda) - F_{m} (0)} \lambda d \lambda \le (\log K^2) \big(F_{m} (K^2) - F_{m} (0)\big). \end{equation} From (\ref{four}), (\ref{five}) and (\ref{six}), we conclude that \begin{equation}\label{seven} \log {\det}'_\pi \Delta_j \ge (\log K^2) \big( F(K^2) - F(0) \big) - \liminf_{m \rightarrow \infty}(\log K^2) \big( F_{m} (K^2) - F_{m} (0) \big). \end{equation} Now Theorem 2.1 yields $$ F (\lambda) = \lim_{\epsilon \rightarrow 0^+} \liminf_{m \rightarrow \infty} F_{m} (\lambda + \epsilon) $$ and $$ F (0) = \lim_{m \rightarrow \infty} F_{m} (0). $$ The last two equalities combined with (\ref{seven}) imply that $\log {\det}'_\pi \Delta_j \ge 0$. Since this is true for all $j=0,1,\ldots,\dim Y$, $Y$ is of determinant class. \end{proof}
proofpile-arXiv_065-494
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Topological invariants in general metric-affine space} \renewcommand{\thesection}{\arabic{section}} \markright{Pontryagin and Euler forms and Chern-Simons...} \setcounter{equation}{0} In this section we consider a metric-affine space ($L_{4}$,$g$) that is a connected 4-dimensional oriented differentiable manifold ${\cal M}$ equipped with a linear connection $\Gamma$ and a metric $g$ of index 1.$^{12}$ We shall use an anholonomic local vector frame $e_{a}$ ($a=1,2,3,4$) and a 1-form coframe $\theta^{a}$ with $e_{a}\rfloor\theta^{b} =\delta^{b}_{a}$ ($\rfloor$ means the interior product). The vector basis $e_{a}$ can be chosen to be pseudo-orthonormal with respect to a metric \begin{equation} g=g_{ab}\theta^{a}\otimes\theta^{b}\;.\label{eq:0} \end{equation} In this case one gets, \begin{equation} g_{ab} := g(e_{a},e_{b}) = diag(+1,+1,+1,-1)\;. \label{eq:1} \end{equation} \par In ($L_{4}$,$g$) a metric $g$ and a connection $\Gamma$ are not compatible in the sense that the $GL(4,R)$-covariant exterior differential (${\cal D}:=d + \Gamma\wedge\ldots$) of the metric does not vanish, \begin{equation} {\cal D}g_{ab} = dg_{ab} - \Gamma^{c}\!_{a}g_{cb} - \Gamma^{c}\!_{b}g_{ac} =: - Q_{ab}\; , \label{eq:2} \end{equation} where $\Gamma^{a}\!_{b}$ is a connection 1-form and $Q_{ab}$ is a nonmetricity 1-form. \par A curvature 2-form $\Omega^{a}\!_{b}$ and a torsion 2-form ${\cal T}^{a}$, \begin{equation} \Omega^{a}\!_{b}=\frac{1}{2}R^{a}\!_{bcd}\theta^{c}\wedge\theta^{d}\;, \qquad {\cal T}^{a}=\frac{1}{2}T^{a}\!_{bc}\theta^{b}\wedge\theta^{c}\;, \label{eq:3} \end{equation} are defined by virtue of the Cartan's structure equations, \begin{eqnarray} \Omega^{a}\!_{b}=d\Gamma^{a}\!_{b}+\Gamma^{a}\!_{c}\wedge\Gamma^{c}\!_{b}\;, \label{eq:4}\\ {\cal T}^{a}={\cal D}\theta^{a}=d\theta^{a}+\Gamma^{a}\!_{b} \wedge\theta^{b}\;. \label{eq:5} \end{eqnarray} \par Let us consider the 4-form \begin{equation} \Pi= B^{b}\!_{a}\!^{q}\!_{p}\Omega^{a}\!_{b}\wedge\Omega^{p}\!_{q} \;, \label{eq:6} \end{equation} where $B^{b}\!_{a}\!^{q}\!_{p}$ is an unknown $GL(4,R)$-invariant tensor. From the form of (\ref{eq:6}) it is easy to get the following symmetry property of this tensor, \begin{equation} B^{b}\!_{a}\!^{q}\!_{p}=B^{q}\!_{p}\!^{b}\!_{a}\;. \label{eq:06} \end{equation} The 4-form (\ref{eq:6}) is proportional to the volume 4-form $\eta$ of the 4-dimensional manifold ${\cal M}$, where \begin{equation} \eta = \frac{1}{4!}\eta_{abcd}\theta^{a}\wedge\theta^{b}\wedge \theta^{c} \wedge\theta^{d}\;, \quad \eta_{abcd}=\sqrt{-det \Vert g_{kl}\Vert}\, \epsilon_{abcd}\;. \end{equation} Here $\epsilon_{abcd}$ is the components of the totaly antisymmetric $GL(4,R)$-invariant Levi-Civita 4-form density $^{12}$ ($\epsilon_{1234}=-1$). \par Since ${\cal D}\eta = d\eta = 0$ as a 5-form on the 4-dimensional manifold ${\cal M}$, one has the identity$^{13}$ \begin{equation} {\cal D}\eta_{abcd} = -\frac{1}{2} Q \eta_{abcd}\;, \qquad Q:=g^{pq} Q_{pq} \;. \label{eq:290} \end{equation} \par The explicit form of the tensor (\ref{eq:06}) should be determined on the basis of the condition that the integral \begin{equation} \int_{{\cal M}}\Pi \label{eq:7} \end{equation} over the oriented 4-demensional manifold {\cal M} without boundary does not depend on the choise of a metric and a connection and therefore the variation of the integrand of (\ref{eq:7}) with respect to a metric and a connection should be equal to an exact form. Here we consider the manifold ${\cal M}$ without boundary for the simplicity. For the manifolds with boundary some additional surface terms should be taken into account.$^{11}$ \par As a consequence of (\ref{eq:0}) the variation with respect to a metric $g$ is determined only by variations of 1-forms $\theta_{a}$ because of the fact that the variation $\delta g_{ab}=0$ when one chooses the pseudo-orthonormal basis $e_{a}$ and gets the condition (\ref{eq:1}). The tensor $B^{b}\!_{a}\!^{q}\!_{p}$ in (\ref{eq:6}) also should not to be varied when the local vector basis $e_{a}$ is chosen to be anholonomic and pseudo-orthonormal because it can be constructed (as an $GL(4,R)$-invariant tensor) only from the metric tensor, Kronecker delta $\delta^{b}_{a}$ and the $GL(4,R)$-invariant totaly antisymmetric Levi-Civita density $\epsilon_{abpq}$. \par The variation of (\ref{eq:6}) yields the expression, \begin{equation} \delta\Pi=2\delta\Gamma^{a}\!_{b}\wedge ({\cal D} B^{b}\!_{a}\!^{q}\!_{p}) \wedge\Omega^{p}\!_{q}+d(2\delta\Gamma^{a}\!_{b}\wedge B^{b}\!_{a}\!^{q}\!_{p} \Omega^{p}\!_{q})\;.\label{eq:8} \end{equation} Here the following relation has been used, \begin{equation} \delta\Omega^{a}\!_{b}\wedge\Phi^{b}\!_{a}=d(\delta\Gamma^{a}\!_{b}\wedge \Phi^{b}\!_{a}) + \delta\Gamma^{a}\!_{b}\wedge {\cal D}\Phi^{b}\!_{a}\; , \label{eq:9} \end{equation} that valids for an arbitrary 2-form $\Phi^{a}\!{_b}$. \par One can see that the variation (\ref{eq:8}) is equal to an exact form, if the tensor $B^{b}\!_{a}\!^{q}\!_{p}$ satisfies the condition, \begin{equation} {\cal D} B^{b}\!_{a}\!^{q}\!_{p}= 0 \;. \label{eq:11} \end{equation} \par In a general metric-affine space ($L_{4}$,$g$) there are only two possibilities to satisfy (up to constant factors) the condition (\ref{eq:11}), \begin{equation} (a)\quad B^{b}\!_{a}\!^{q}\!_{p} =\delta^{b}_{a}\delta^{q}_{p}\;, \qquad (b)\quad B^{b}\!_{a}\!^{q}\!_{p} =\delta^{b}_{p}\delta^{q}_{a}\;.\label{eq:12} \end{equation} \par In the case (a) the 4-form $\Pi$ (\ref{eq:6}) reads, \begin{equation} \Pi_{\Omega}=\Omega^{a}\!_{b}\wedge\Omega^{b}\!_{a} = \mbox{\rm Tr}(\Omega \wedge\Omega )\;, \label{eq:13} \end{equation} and in the case (b) one has, \begin{equation} \Pi_{tr\Omega}=\Omega^{a}\!_{a}\wedge \Omega^{b}\!_{b}= \mbox{\rm Tr}\Omega \wedge\mbox{\rm Tr}\Omega \;. \label{eq:15} \end{equation} We see that the 4-forms (\ref{eq:13}) and (\ref{eq:15}) are equal up to constant factors to the well-known Pontryagin forms.$^{11,12}$ \renewcommand{\thesection}{\Roman{section}.} \section{The topological invariants in a Weyl- \newline Cartan space} \renewcommand{\thesection}{\arabic{section}} \markright{Pontryagin and Euler forms and Chern-Simons...} \setcounter{equation}{0} A Weyl-Cartan space $Y_{4}$ is a space with a metric, curvature, torsion and nonmetricity which obeys the constraint, \begin{equation} Q_{ab} = \frac{1}{4}g_{ab}Q\;. \label{eq:21} \end{equation} This constraint can be introduced into the variational approach with the help of the method of Lagrange multipliers. In this case the integral (\ref{eq:7}) has to be modified, \begin{equation} \int_{{\cal M}}\left (\Pi + \Lambda^{ab}\wedge (Q_{ab} - \frac{1}{4}g_{ab}Q) \right )\;, \label{eq:22} \end{equation} where the Lagrange multiplier $\Lambda^{ab}$ is a tensor-valued 3-form with the properties, \begin{equation} \Lambda^{ab}=\Lambda^{ba}\;, \qquad \Lambda^{a}\!_{a}=0 \;. \label{eq:220} \end{equation} \par The variation of (\ref{eq:22}) with respect to $\theta^{a}$, $\Gamma^{a}\!_{b}$ and the Lagrange mutiplier yields that the following variational derivatives have to vanish identically, \begin{eqnarray} \delta \Gamma^{a}\!_{b}\;: &{\cal D}(B^{b}\!_{a}\!^{q}\!_{p})\wedge \Omega^{p}\!_{q} - \Lambda^{b}\!_{a} = 0 \;, & \label{eq:23}\\ \delta\Lambda^{ab}\;: &Q_{ab} - \frac{1}{4}Q g_{ab} = 0 \;. &\label{eq:24} \end{eqnarray} As in the previous section the variational derivative with respect to $\theta^{a}$ is absent because of the fact that there is no an explicit dependence on $\theta^{a}$ of the integrand expression in (\ref{eq:22}). \par The identity (\ref{eq:23}) in $Y_{4}$ is equivalent to the following identities, \begin{eqnarray} &\Lambda^{ba}= ({\cal D} - \frac{1}{4}Q)B^{(ba)q}\!_{p} \wedge \Omega^{p}\!_{q}\; , \label{eq:25}\\ &({\cal D} - \frac{1}{4}Q)B^{[ba]q}\!_{p} \wedge \Omega^{p}\!_{q}= 0\; , \label{eq:26}\\ &{\cal D}B^{a}\!_{a}\!^{q}\!_{p} \wedge\Omega^{p}\!_{q} = 0\;. \label{eq:27} \end{eqnarray} \par The identities (\ref{eq:26}), (\ref{eq:27}) are satisfied in the following four cases: \begin{eqnarray} &(a)\quad B^{baq}\!_{a}\!^{q}\!_{p} =g^{ba}\delta^{q}_{p}\;, \qquad (b)\quad B^{baq}\!_{p} =\delta^{b}_{p}g^{qa}\;, \label{eq:28}\\ &(c)\quad B^{baq}\!_{p} =g^{bq}\delta^{a}_{p}\;, \qquad (d)\quad B^{baq}\!_{p} =\eta^{baq}\!_{p}\;.\label{eq:29} \end{eqnarray} In the case (d) one has to use (\ref{eq:2}), (\ref{eq:21}) and (\ref{eq:290}). \par The equality (\ref{eq:25}) determines the Lagrange miltiplier. In all cases (a)-(d) one has $\Lambda^{ab}=0$. This means that the Weyl-Cartan constraint (\ref{eq:24}) can be imposed both before and after the variational procedure. \par The cases (a) and (b) coinside with (\ref{eq:12}) and yield for a Weyl-Cartan space $Y_{4}$ the Pontryagin forms (\ref{eq:13}) and (\ref{eq:15}) of the previous section. The cases (c) and (d) appear in $Y_{4}$ but not in ($L_{4}$,$g$). \par In the case (c) one has the Pontryagin form, \begin{equation} \Pi_{CW}=\Omega^{ab}\wedge \Omega_{ab} = \mbox{\rm Tr}(\Omega \wedge \Omega ^{T})\;,\label{eq:30} \end{equation} where $\Omega^{T}$ means the transpose of $\Omega$. In $Y_{4}$ with the help of the relation, \begin{equation} \Omega_{ab}= \Omega_{[ab]} + \frac{1}{4}g_{ab}\mbox{\rm Tr}\Omega\;, \label{eq:31} \end{equation} (\ref{eq:30}) can be decomposed as follows, \begin{equation} \Omega^{ab}\wedge \Omega_{ab}=\Omega^{[ab]}\wedge \Omega_{[ab]} + \frac{1}{4} \mbox{\rm Tr}\Omega \wedge \mbox{\rm Tr}\Omega\;. \label{eq:32} \end{equation} On the other hand the Pontryagin form (\ref{eq:13}) in $Y_{4}$ has the decomposition, \begin{equation} \Omega^{a}\!_{b}\wedge \Omega^{b}\!_{a}= -\Omega^{[ab]}\wedge\Omega_{[ab]} +\frac{1}{4}\mbox{\rm Tr}\Omega\wedge \mbox{\rm Tr}\Omega \;. \label{eq:33} \end{equation} Therefore in a Weyl-Cartan space $Y_{4}$ one has two fundamental Pontryagin forms, which are equal up to constant factors to \begin{equation} \Pi_{C}=\Omega^{[ab]}\wedge \Omega_{[ab]}\;, \qquad \Pi_{W}=\mbox{\rm Tr}\Omega\wedge \mbox{\rm Tr}\Omega \;. \label{eq:34} \end{equation} The former form is the volume preserving Pontryagin form and the latter one is the dilatonic Pontryagin form. \par In the case (d) we get the Euler form in a Weyl-Cartan space $Y_{4}$, \begin{equation} {\cal E}= \eta^{b}\!_{a}\!^{q}\!_{p}\Omega^{a}\!_{b}\wedge \Omega^{p}\!_{q} \;. \label{eq:35} \end{equation} One can use the holonomic coordinate basis $e_{\alpha} = \partial_{\alpha}$ and express the topological invariant corresponding to (\ref{eq:35}) in the component form, \begin{equation} \int_{{\cal M}}{\cal E} = \int_{{\cal M}}E \sqrt{-g}dx^{1}\wedge dx^{2} \wedge dx^{3}\wedge dx^{4}\;, \label{eq:36} \end{equation} \begin{equation} E = R^{2} - (R_{\alpha\beta} +\tilde{R}_{\alpha\beta})(R^{\beta\alpha} + \tilde{R}^{\beta\alpha}) + R_{\alpha\beta\mu\nu}R^{\mu\nu\alpha\beta}\; , \label{eq:37} \end{equation} where $R^{\alpha}\!_{\beta\mu\nu}$ are the components of the curvature 2-form in a holonomic basis, the following notations being used, $R_{\alpha\beta} = R^{\sigma}\!_{\alpha\sigma\beta}$, $\tilde{R}_{\alpha\beta} = R_{\alpha\sigma\beta}\!^{\sigma}$, $R = R_{\sigma}\!^{\sigma}$. \par The Gauss-Bonnet-Chern Teorem$^{10}$ states the relation of the integral (\ref{eq:36}) over the oriented compact manifold ${\cal M}$ without boundary with the Euler characteristic of this manifold. The explicit proof using a holonomic basis of the independence of (\ref{eq:36}) on the choice of a metric and a connection of a Weyl-Cartan space $Y_{4}$ is explained in Ref. 14. \renewcommand{\thesection}{\Roman{section}.} \section{Chern-Simons terms in a Weyl- \newline Cartan space} \renewcommand{\thesection}{\arabic{section}} \markright{Pontryagin and Euler forms and Chern-Simons...} \setcounter{equation}{0} It is well known that in ($L_{4}$,$g$) the Pontryagin forms can be represented as the exterior derivatives of the $GL(4,R)$ Chern-Simons terms,$^{12}$ \begin{eqnarray} &\Pi_{\Omega}=d{\cal C}_{\Omega}\;, \qquad {\cal C}_{\Omega}=\Gamma^{b}\!_{a} \wedge \Omega^{a}\!_{b} - \frac{1}{3}\Gamma^{b}\!_{a}\wedge\Gamma^{a}\!_{c} \wedge \Gamma^{c}\!_{b}\;, \label{eq:40}\\ &\Pi_{W}=d{\cal C}_{W}\;, \qquad {\cal C}_{W} = \frac{1}{2}Q\wedge \Omega^{a} \!_{a}\;. \label{eq:41} \end{eqnarray} \par It is easy to see that Pontryagin form $\Pi_{C}$ (\ref{eq:34}) in a Weyl-Cartan space $Y_{4}$ can be represented in an analogous manner, \begin{equation} \Pi_{C} = d{\cal C}_{C}\;, \qquad {\cal C}_{C}= \Gamma^{[b}\!_{a]}\wedge \Omega^{[a}\!_{b]} - \frac{1}{3}\Gamma^{[b}\!_{a]} \wedge\Gamma^{[a}\!_{c]}\wedge \Gamma^{[c}\!_{b]}\;. \label{eq:42} \end{equation} \par As it was pointed out in Ref. 12, the Euler form (\ref{eq:35}) in the framework of a Riemann-Cartan space can be expressed in terms of the corresponding Chern-Simons type construction, \begin{equation} {\cal E} = d{\cal C}_{{\cal E}}\;, \qquad {\cal C}_{{\cal E}} = \eta^{b}\!_{a}\!^{q}\!_{p} \left (\Omega^{a}\!_{b}\wedge \Gamma^{p}\!_{q} - \frac{1}{3}\Gamma^{a}\!_{b}\wedge \Gamma^{p}\!_{f}\wedge\Gamma^{f}\!_{q} \right ) \;. \label{eq:43} \end{equation} \par Let us prove that formula (\ref {eq:43}) is also valid in a Weyl-Cartan space $Y_{4}$. The proof is based on the two Lemmas. \par {\it Lemma 1}. If the equality \begin{equation} {\cal D}\eta^{b}\!_{a}\!^{q}\!_{p}=0\;, \label{eq:44} \end{equation} is valid, then the identity (\ref{eq:43}) is fulfilled. \par {\it Proof}. In anholonomic orthonormal frames one has $d\eta^{b}\!_{a}\!^{q}\!_{p}=0$, and therefore (\ref{eq:44}) yields, \begin{equation} \Gamma^{b}\!_{f}\eta^{f}\!_{a}\!^{q}\!_{p}-\Gamma^{f}\!_{a}\eta^{b}\!_{f}\! ^{q}\!_{p}+\Gamma^{q}\!_{f}\eta^{b}\!_{a}\!^{f}\!_{p}-\Gamma^{f}\!_{p} \eta^{b}\!_{a}\!^{q}\!_{f}=0\;. \label{eq:45} \end{equation} After multiplying (\ref{eq:45}) externally by the 3-form $\Gamma^{a}\!_{s} \wedge\Gamma^{s}\!_{b}\wedge\Gamma^{p}\!_{q}\wedge$, one gets the $Y_{4}$-identity, \begin{equation} \eta^{b}\!_{a}\!^{q}\!_{p}\Gamma^{a}\!_{s}\wedge\Gamma^{s}\!_{b}\wedge \Gamma^{p}\!_{f}\wedge\Gamma^{f}\!_{q}=0\;. \label{eq:46} \end{equation} After multiplying (\ref{eq:45}) externally by the 3-form $\Omega^{a}\!_{b} \wedge\Gamma^{p}\!_{q}\wedge$, one gets the second $Y_{4}$-identity, \begin{equation} \eta^{b}\!_{a}\!^{q}\!_{p}(2\Omega^{a}\!_{b}\wedge\Gamma^{p}\!_{f}\wedge \Gamma^{f}\!_{q}-\Omega^{a}\!_{f}\wedge\Gamma^{f}\!_{b}\wedge\Gamma^{p}\!_{q} +\Gamma^{a}\!_{f}\wedge\Omega^{f}\!_{b}\wedge\Gamma^{p}\!_{q})=0\;. \label{eq:47} \end{equation} Now using the identities (\ref{eq:46}) and (\ref{eq:47}), the Cartan's structure equation (\ref{eq:4}) and the Bianchi identity, \begin{equation} {\cal D}\Omega^{a}\!_{b}=d\Omega^{a}\!_{b}+\Gamma^{a}\!_{f}\wedge \Omega^{f}\!_{b}-\Omega^{a}\!_{f}\wedge\Gamma^{f}\!_{b}=0\;, \end{equation} let us perform the exterior differentiation of the Chern-Simons term ${\cal C}_{{\cal E}}$ (\ref{eq:43}) and get, \begin{eqnarray} &d{\cal C}_{{\cal E}}-{\cal E}=\frac{1}{3}\eta^{b}\!_{a}\!^{q}\!_{p} \Gamma^{a}\!_{s}\wedge\Gamma^{s}\!_{b}\wedge\Gamma^{p}\!_{f}\wedge \Gamma^{f}\!_{q}\nonumber \\ &-\frac{2}{3}\eta^{b}\!_{a}\!^{q}\!_{p} (2\Omega^{a}\!_{b}\wedge\Gamma^{p}\!_{f}\wedge\Gamma^{f}\!_{q} -\Omega^{a}\!_{f}\wedge\Gamma^{f}\!_{b}\wedge\Gamma^{p}\!_{q} +\Gamma^{a}\!_{f}\wedge\Omega^{f}\!_{b}\wedge\Gamma^{p}\!_{q})=0\;, \end{eqnarray} as was to be proved. \par {\it Lemma 2}. The equality (\ref{eq:44}) is valid if and only if the space under consideration is a Weyl-Cartan space $Y_{4}$. \par {\it Proof}. In a general ($L_{4}$,$g$) space one has, \begin{equation} {\cal D}\eta^{b}\!_{a}\!^{q}\!_{p}=\eta^{q}\!_{map}\tilde{Q}^{bm} - \eta^{b}\!_{map}\tilde{Q}^{qm}\;, \label{eq:48} \end{equation} where $\tilde{Q}^{bm} := Q^{bm}- \frac{1}{4}g^{bm}{Q}$ is the tracefree part of the nonmetricity 1-form, $\tilde{Q}^{b}\!_{b} =0$. For a Weyl-Cartan space $Y_{4}$ one has $\tilde{Q}^{bm}=0$ and the sufficient condition of the Lemma is evident. The necessary condition of the Lemma is the consequence of the fact that the vanishing of (\ref{eq:48}) leads to the equality, \begin{equation} g^{ab}\tilde{Q}^{pq}-g^{bp}\tilde{Q}^{aq}-g^{aq} \tilde{Q}^{bp}+g^{pq}\tilde{Q}^{ab}=0\;, \end{equation} which yields $\tilde{Q}^{pq}=0$, as was to be proved. \renewcommand{\thesection}{\Roman{section}.} \section{Conclusions} \renewcommand{\thesection}{\arabic{section}} \markright{Pontryagin and Euler forms and Chern-Simons...} We have proved the existence of the Pontryagin and Euler forms in a Weyl-Cartan space on the basis of the variational method with Lagrange multipliers. It has been discovered that the Pontryagin form, $\Pi_{C}=\Omega^{[ab]}\wedge \Omega_{[ab]}$, and Euler form, ${\cal E}= \eta^{b}\!_{a}\!^{q}\!_{p}\Omega^{a}\!_{b}\wedge \Omega^{p}\!_{q}$, which are specific for a Riemann-Cartan space, also exist in a Weyl-Cartan space. With the help of these forms the topological invariants of a Weyl-Cartan space which do not depend on the choice of a metric and a connection are constructed. It has been proved that these forms can be expressed via the exterior derivatives of the corresponding Chern-Simons terms in a Weyl-Cartan space (see (\ref{eq:42}) and (\ref{eq:43}), respectively). From the Lemma 2 proved it follows that the relation (\ref{eq:43}) is not valid in the more general geometry than the Weyl-Cartan one. \newpage \vskip 0.6cm \begin{description} \item{$^{1}$} M.B. Green,J.H. Schwarz and E. Witten, {\em Superstring Theory}, 2 volumes (Cambridge University Press, Cambridge, 1987). \item{$^{2}$} G. 't Hooft, M. Veltman, {\em Ann. Inst. H. Poincar\'{e}} {\bf 20}, 69 (1974). \item{$^{3}$} K.S. Stelle, {\em Phys. Rev.} {\bf D16}, 953 (1977). \item{$^{4}$} R. Bach, {\em Math. Z.} {\bf 9}, 110 (1921). \item{$^{5}$} C. Lanczos, {\em Ann. Math.(N.Y.)} {\bf 39}, 842 (1938). \item{$^{6}$} J.R. Ray, {\em J. Math. Phys.} {\bf 19}, 100 (1978). \item{$^{7}$} V.N. Tunjak, {\em Izvestija vyssh. uch. zaved. (Fizika)} N9, 74 (1979) [in Russian]. \item{$^{8}$} H.T. Nieh, {\em J. Math. Phys.} {\bf 21}, 1439 (1980). \item{$^{9}$} K. Hayashi, T. Shirafuji, {\em Prog. Theor. Phys.} {\bf 65}, 525 (1981). \item{$^{10}$} R. Sulanke und P. Wintgen, {\em Differentialgeometrie und faserb\"undel} (Hoch\-schulb\"ucher f\"ur mathematik, band 75)(VEB Deutscher Verlag der Wis\-senshaften, Berlin, 1972). \item{$^{11}$} T. Eguchi, P.B. Gilkey and A.J. Hanson, {\em Phys. Reports} {\bf 66}, 213 (1980). \item{$^{12}$} F.W. Hehl, J.D. McCrea, E.W. Mielke and Yu. Ne'eman, {\em Phys. Reports} {\bf 258}, 1 (1995). \item{$^{13}$} R. Tresguerres, {\em J. Math. Phys.} {\bf 33}, 4231 (1992). \item{$^{14}$} O.V. Babourova, B.N. Frolov, {\em Gauss-Bonnet type identity in Weyl-Cartan space-time}, LANL e-archive gr-qc/9608... (1996). \end{description} \end{document}
proofpile-arXiv_065-495
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction}\label{sec:intr} A typical quantity to analyze the nature of the perturbative expansion in Quantum Field Theory is the partition function \begin{equation} Z ( \lambda ) = {1 \over Z_0} \int \left[ {\rm d} \phi \right] e^{- S \left[ \phi \right]} \label{eq:partfunc} \end{equation} with \begin{equation} S \left[ \phi \right] = \int {\rm d^d} x \left[ {1 \over 2} \left( \partial_{\mu} \phi \right)^2 + {1 \over 2} m^2 \phi^2 + {\lambda \over 4} \phi^4 \right]. \label{eq:action} \end{equation} The normalization factor $1/Z_0$ is the partition function of the free field ($Z \rightarrow 1 $ when $ \lambda \rightarrow 0 $). The analysis of the perturbative expansion of any Green's function goes along similar lines as in the case of $Z$. In the example above we consider a scalar field theory for simplicity. The traditional argument for understanding the divergent nature of the perturbative expansion can be traced back to Dyson~\cite{dy}. Although the form was different, the content of his argument is captured by the following statement: ``If the perturbative series were to converge to the exact result, the function being expanded would be analytic in $\lambda$ at $\lambda = 0$. But the function (Z for example) is not analytic in $\lambda$ at that value. Therefore, as a function of $\lambda$, the perturbative series is either divergent or converges to the wrong answer." Estimations of the large order behavior of the coefficients of the perturbative series showed that the first possibility is the one actually realized \cite{lo,bw}. That Z, as a function of $\lambda$, is not analytic at $\lambda = 0$, can be guessed by simply noting that if in its functional integral representation (Eq.~(\ref{eq:partfunc})) we make the real part of $\lambda$ negative, the integral diverges. In fact, there is a branch cut in the first Riemann sheet that can be chosen to lie along the negative real axis, extending from $\lambda = - \infty$ to $ \lambda = 0$ \cite{bw0,bs}. The above argument is very powerful and extends to the perturbative series of almost all other nontrivial field theories. It has also motivated a series of very important calculations of the large order behavior of the perturbative coefficients \cite{lo}, general analysis on the structure of field theories \cite{tHooft}, as well as improvements over perturbative computations of different physical quantities \cite{lo}. For all its power, it is fair to say that the argument, as almost any other {\it reductio ad absurdum} type of argument, fails to point towards a solution of the problem of divergence. It is only through the indirect formalism of Borel transforms that questions of recovery of the full theory from its perturbative series can be discussed \cite{bs,gh,zinn}. In this paper an alternative way of understanding the divergent nature of the perturbative series is presented. This way of understanding the problem complements the traditional argument briefly described above, hopefully illuminating aspects that the traditional approach leaves obscure. In particular, as we will see, the arguments in this paper point directly towards the aspects of the perturbative series that need to be modified to achieve a convergent series. It is hoped that the way of understanding the problem presented here will help to provide new insights into the urgent problem of extracting non-perturbative information out of Quantum Field Theories. In section~\ref{sec:leb} we develop our analysis of the divergence of perturbation theory. In Sec.~\ref{sec:converg} we point out the ingredients that, according to the analysis of Sec.~\ref{sec:leb}, a modification of perturbation theory would need to achieve convergence. We also present a remarkable formula~(\ref{win}) that allows us to implement such modifications in terms of Gaussian integrals, paving the way to the application of this convergent modified perturbative series to Quantum Field Theories. The proof of the properties of the function~(\ref{win}) is done in Appendix 1. In section~\ref{Improvement} we analyze recent work on the convergence of various optimized expansions~[12-19] in terms of the ideas presented here. In Sec.~\ref{sec:concl} we summarize our results and mention directions of the work currently in preparation. Finally, in Appendix 2, we apply the ideas of this paper in a simple but illuminating example for which we actually develop a convergent series by modifying the aspects of the perturbative series pointed out by our analysis as the source of divergence. \section{Lebesgue's Dominated Convergence Theorem and Perturbation Theory}\label{sec:leb} \subsection{The wrong step in perturbation theory}\label{sec:wrong} Although the notation will not always be explicit, we work in an Euclidean space of dimension smaller than 4 and in a finite volume. Let's remember how the perturbative series is generated in the functional integral formalism for a quantity like $Z$: \begin{eqnarray}\label{eq:pertZ} Z ( \lambda ) & = & \int \left[ {\rm d} \phi \right] e^{- \int {\rm d^d} x \left[ {1 \over 2} \left( \partial_{\mu} \phi \right)^2 + {1 \over 2} m^2 \phi^2 \right] - {\lambda \over 4} \int {\rm d^d} x \phi^4 } \\ \label{eq:first} & = & \int \left[ {\rm d} \phi \right] \sum_{n=0}^{\infty} {\left( -1 \right)^n \over n!} \left( {\lambda \over 4} \int {\rm d^d} x \phi^4 \right)^n e^{- \int {\rm d^d} x \left[ {1 \over 2} \left( \partial_{\mu} \phi \right)^2 + {1 \over 2} m^2 \phi^2 \right] } \\ \label{eq:second} & = & \sum_{n=0}^{\infty} \int \left[ {\rm d} \phi \right] {\left( -1 \right)^n \over n!} \left( {\lambda \over 4} \int {\rm d^d} x \phi^4 \right)^n e^{- \int {\rm d^d} x \left[ {1 \over 2} \left( \partial_{\mu} \phi \right)^2 + {1 \over 2} m^2 \phi^2 \right] } \end{eqnarray} The final sum is in practice truncated at some finite order $N$. The functional integrals that give the contribution of every order $n$ are calculated using Wick's theorem and Feynman's diagram techniques with the corresponding renormalization. We see then that the generation of the perturbative series in the functional integral formalism is a two step process. First (\ref{eq:first}) the integrand is expanded in powers of the coupling constant, and then (\ref{eq:second}) the sum is interchanged with the integral\footnote{ In this paper we will often use the familiar word ``integrand" to refer to $e^{-S}$ or any functional inside the functional integration symbol. It would be more precise to preserve this word for $e^{-S_{\rm Int}}$ in the measure defined by the free field. The terminology used here is, however, common practice in the Quantum Field Theory literature and also helps to emphasize the similarities with the intuitive finite dimensional case presented below. } It will be convenient to have a simpler example in which the arguments of this paper become very transparent. Consider the simple integral \begin{eqnarray}\label{eq:pertsimple} z ( \lambda ) & = & { 1 \over \sqrt{\pi} } \int_{- \infty}^{\infty} {\rm d} x e^{- \left( x^2 + {\lambda \over 4} x^4 \right)} \end{eqnarray} and its corresponding perturbative expansion \begin{eqnarray}\label{eq:pertsimple1} z ( \lambda ) & = & { 1 \over \sqrt{\pi} } \int_{- \infty}^{\infty} {\rm d} x \sum_{n=0}^{\infty} {\left( -1 \right)^n \over n! } \left( {\lambda \over 4} x^4 \right)^n e^{- x^2 } \\ \label{eq:pertsimple2} & = & { 1 \over \sqrt{\pi} } \sum_{n=0}^{\infty} \int_{- \infty}^{\infty} {\rm d} x {\left( -1 \right)^n \over n! } \left( {\lambda \over 4} x^4 \right)^n e^{- x^2 } \\ \label{eq:simplecoef} & \equiv & \sum_{n=0}^{\infty} \left( -1 \right)^n c_n \ \lambda^n. \end{eqnarray} This simple integral has been used many times in the past as a paradigmatic example of the divergence of perturbation theory \cite{zinn}. It is then specially suited for a comparison between the traditional arguments and the ones presented in this paper. Again we see the two step process to generate the perturbative series. First the integrand is expanded in powers of $\lambda$~(\ref{eq:pertsimple1}) and then the sum is interchanged with the integral~(\ref{eq:pertsimple2}) . In this simple example the perturbative coefficients can be calculated exactly for arbitrary $n$. In the large $n$ limit they become: \begin{equation}\label{eq:simplecoeflargen} c_n \sim {\sqrt{2} \over 2 \pi } \left( n -1 \right)! \quad {\rm when} \quad n \rightarrow \infty. \end{equation} With such factorial behavior, the series diverges for all $ \lambda$ different from zero as is well known. On the other hand the function $z ( \lambda )$, as defined in Eq.~(\ref{eq:pertsimple}), gives a well defined positive real number for every positive real $\lambda$. Therefore one or both of the two steps done to generate the perturbative series must be wrong. Similarly, in the functional integral case normalized with respect to the free field (\ref{eq:partfunc}), $Z$ is a well defined number while its perturbative series diverges. One or both of the two steps must be wrong. The first step, the expansion of the integrand in powers of $ \lambda$, is clearly correct. As the integrand (not the integral!) is analytic in $ \lambda$ for every finite $ \lambda$, the expansion merely corresponds to a Taylor series. The second step, the interchange of sum and integral, must therefore be the wrong one. The next obvious step is then to recall the theorems that govern the interchange between sums and integrals, to understand in detail why this is wrong in our case. The most powerful theorem in this respect is the well known theorem of Dominated Convergence of Lebesgue. In a simplified version, enough for our purposes, it says the following: \begin{quote} Let $f_N$ be a sequence of integrable functions that converge pointwisely to a function $f$ \begin{equation}\label{eq:pointconv} f_N \longrightarrow f \quad {\rm as} \quad N \rightarrow \infty \end{equation} and bounded in absolute value by a positive integrable function $h$ (dominated) \begin{equation}\label{eq:bound} |f_N| \leq h \ ,\quad \forall N. \end{equation} Then, it is true that \begin{equation}\label{eq:thesis} \lim_{N \rightarrow \infty} \int f_N = \int \lim_{N \rightarrow \infty} f_N = \int f . \end{equation} \end{quote} As a special case, if the convergence~(\ref{eq:pointconv}) is uniform and the measure of integration is finite, then the interchange is also valid. It should be emphasized that Lebesgue's theorem follows from the axioms of abstract measure theory. Therefore if the problem under consideration involves a well defined measure, as is the case for the Quantum Field Theories considered here~\cite{glja}, the theorem holds. In our case we can write formally\footnote{See previous footnote.}, \begin{equation}\label{eq:fNfunctint} f_N \left[ \phi (x) \right]= {1 \over Z_0} \sum_{n=0}^{N} {\left( -1 \right)^n \over n!} \left( {\lambda \over 4} \int {\rm d^d} x \phi^4 \right)^n e^{- \int {\rm d^d} x \left[ {1 \over 2} \left( \partial_{\mu} \phi \right)^2 + {1 \over 2} m^2 \phi^2 \right] } \end{equation} for the functional integral case, and \begin{equation}\label{eq:fNsimpltint} f_N (x) = { 1 \over \sqrt{\pi} } \sum_{n=0}^{N} {\left( -1 \right)^n \over n!} \left( {\lambda \over 4} x^4 \right)^n e^{- x^2 } \end{equation} for the simple integral example. One important aspect of the dominated convergence theorem approach to analyze the divergence of perturbation theory is that it focuses on the integrands, objects relatively simple to analyze. On the contrary, the analyticity approach briefly described in the introduction focuses on the integrals, that are much more difficult to analyze. So, before we try to understand what aspects of the dominated convergence theorem fail in our case, let's see the ``phenomena" (the integrand) for the intuitive simple example. \begin{figure}[htp] \hbox to \hsize{\hss\psfig{figure=ex00204.eps,width=0.9\hsize}\hss} \caption{Exact integrand, ${\rm zero}^{{\rm th}}$, second and fourth perturbative approximations. $\lambda = 1$.} \label{fig1} \end{figure} In figure ~\ref{fig1}, the exact integrand, together with some perturbative approximations, are displayed. We can appreciate the way in which the successive approximations behave. For small $x$, and up to some critical value that we call $x_{c,N}$, where the subindex $c$ stands for {\it critical} while the subindex $N$ indicates that this value changes with the order, the perturbative integrands approximate very well the exact integrand. Even more, $x_{c,N}$ grows with $N$. But for $x$ bigger that $x_{c,N}$ a ``bump'' begins to emerge. The height of these bumps, as we will see in detail shortly, grows factorially with the order, while the width remains approximately constant. So, the larger the order in perturbation theory, the larger the region in which the perturbative integrands approximate very well the exact integrand, but the stronger the upcoming deviation. As we will see shortly, it is precisely this deviation that is responsible for the divergence of the perturbative series and the famous factorial growth. We will also see that an exactly analogous phenomena happens in the functional integral case and is again the responsible for the divergence of the perturbative series. Returning to the problem of understanding the aspect of the dominated convergence theorem that fail in the perturbative series we will now show that the sequence of integrands of Eq.~(\ref{eq:fNfunctint}) and Eq.~(\ref{eq:fNsimpltint}) converge respectively to the exact integrands \begin{equation}\label{eq:exIntd} F = {1 \over Z_0} \ e^{- \int {\rm d^d} x \left[ {1 \over 2} \left( \partial_{\mu} \phi \right)^2 + {1 \over 2} m^2 \phi^2 \right] - {\lambda \over 4} \int {\rm d^d} x \phi^4 } \end{equation} and \begin{equation}\label{eq:exintd} f = {1 \over \sqrt{\pi}} e^{- \left( x^2 + {\lambda \over 4} x^4 \right)} \end{equation} but {\it not} in a {\it dominated} way. This is, there is no positive integrable function $h$ satisfying the property of Eq.~(\ref{eq:bound}). \subsection{Failure of domination in the simple example}\label{sec:simpleint} That the sequence of integrands of Eq.~(\ref{eq:fNfunctint}) and Eq.~(\ref{eq:fNsimpltint}) converge respectively to the exact integrands (\ref{eq:exIntd}) and (\ref{eq:exintd}) is obvious, since, as mentioned before, for finite $\lambda$ they are analytic functions of $\lambda$ and so their Taylor expansion converge (at least for finite field strength). To see the failure of the domination hypothesis it is convenient to analyze the ``shape" of every term of $f_N$. Namely, for the field theory case, \begin{equation}\label{eq:cnFT} c_n \left[ \phi (x) \right] \equiv {1 \over Z_0} {\left( -1 \right)^n \over n!} {1 \over 4^n} \left( \lambda \int {\rm d^d} x \phi^4 \right)^n e^{- \int {\rm d^d} x \left[ {1 \over 2} \left( \partial_{\mu} \phi \right)^2 + {1 \over 2} m^2 \phi^2 \right] } \end{equation} while for the simple integrand \begin{equation}\label{eq:cnSI} c_n (x) = {1 \over \sqrt{\pi}} {\left( -1 \right)^n \over n!} \left( {\lambda \over 4} \right)^n x^{4n} e^{- x^2 }. \end{equation} In this section we analyze the failure of the domination hypothesis for the simple example~(\ref{eq:pertsimple}) because, as it turns out, it is remarkably similar to the Quantum Field Theory example analyzed in the next section. In Fig.~\ref{fig2} we can inspect the functions $c_3 (x)$ and $c_4(x)$ for $\lambda = 1$ corresponding to the simple integrand case that we analyze first. The maximum of $c_n (x)$ is reached at \begin{equation}\label{eq:xmax} x_{{\rm max}} = \pm (2 n)^{1/2}. \end{equation} There, for large $n$, the function takes the value \begin{equation}\label{eq:cnmax} c_n (x_{{\rm max}}) = {1 \over 2 \pi^{3/2}} (-1)^n (n-1)! \ \lambda^n . \end{equation} On the other hand, the width remains constant as n increases as can be seen by a Gaussian approximation around the maximum $x_{{\rm max}} = (2 n)^{1/2}$: \begin{equation}\label{gaussapp} c_n (x) \approx {1 \over 2 \pi^{3/2}} (-1)^n (n-1)! \ {\rm exp} \left[ -2 (x - (2n)^{1/2})^2 \right] \ \lambda^n . \end{equation} The integration of this Gaussian approximation gives, for large $n$ \begin{equation}\label{gaussapp2} \int {\rm d} x \ c_n (x) \approx {1 \over 2} { \sqrt{2} \over 2 \pi} (-1)^n (n-1)! \ \lambda^n \end{equation} in accordance with Eq.~(\ref{eq:simplecoeflargen}) if we take into account the factor of 2 coming from the two maxima $\pm (2 n)^{1/2}$. \begin{figure}[htp] \hbox to \hsize{\hss\psfig{figure=ex04c3c4.eps,width=0.9\hsize}\hss} \caption{Exact integrand, and fourth perturbative approximation together with the third, and fourth terms. $\lambda = 1$.} \label{fig2} \end{figure} The mechanism of convergence of the $f_N $'s to $f$ becomes clear now. The $f_N $'s are made out of a pure Gaussian (the ``free" term) plus ``bumps" (the perturbative corrections) oscillating in sign (see Fig.~\ref{fig2}). The maxima of these bumps grows factorially with the order while their width remain approximately constant (more specifically, the Gaussian approximation around the maxima (Eq.~(\ref{gaussapp})), that becomes exact when the order goes to infinity, has a variance independent of the order). For fixed $N$, and for $x$ smaller than a certain value, the bumps delicately ``almost" cancel each other, leaving only a small remnant that modifies the free integrand into the interacting one. However, for $x$ larger than that value, the last bump begins to emerge and, being the last, does not have the next one to get cancelled (in the $N \rightarrow \infty$ limit, there is no last bump and the convergence is achieved for every x). Consequently, after a certain value $x_{c,N}$, the function $f_N$ deviates strongly from $f$ and is governed by the uncanceled $N^{{\rm th}}$ bump only, of height proportional to $(N-1)!$ and finite variance. This is so because, since the height of the bump grows factorially with the order, for $N$ large enough the last bump is far greater than all the previous ones, so it remains almost completely uncanceled. Since the variance of the bumps is independent of the order, this means that for every finite order, there is a region of {\it finite measure} in which the perturbative integrand is of the order of the height of the last bump. In figure ~\ref{fig2} we can see how the function $c_4 (x)$ is left almost completely uncanceled by $c_3 (x)$ and dominates the deviation of $f_4$ from $f$. That $x_{c,N}$ (the value of $|x|$ up to which the perturbative integrand very accurately approximate the exact one) grows with $N$, going to infinity when $N \rightarrow \infty$, is a simple consequence of Taylor's theorem applied to the analytic function $e^{- \lambda x^4 /4}$. The above analysis makes clear the failure of domination of the sequence of Eq.~(\ref{eq:fNsimpltint}) towards $f$ (Eq.~(\ref{eq:exintd})). Indeed, any positive function $h(x)$ with the property \begin{equation}\label{eq:domSI} |f_N (x)| \leq h (x)\ ,\quad \forall N \end{equation} fails to be integrable, since it has to ``cover" the bump, whose area grows factorially with $N$. So, although the sequence of $f_N (x)$'s converges to $f(x)$, the convergence is not dominated as we wanted to show. Eq.~(\ref{gaussapp2}), together with the above comments, indicate that the same reason for which the sequence of integrands (\ref{eq:cnSI}) fails to be dominated, is the one that produces the factorial growth in the perturbative series. In the Field Theory case, although we can not rely on figures like~\ref{fig1} and \ref{fig2} to guide our intuition, as we will show now, the analogy with the simple integral example is so close that the interpretation is equally transparent. \subsection{Failure of domination in Quantum Field Theory}\label{sec:QM} For Quantum Field Theory, as for the simple example analyzed above, it is convenient to consider every term $c_n [ \phi (x) ]$ (Eq.~\ref{eq:cnFT}) of the perturbative approximation $f_N$ (Eq.~\ref{eq:fNfunctint}) of the exact integrand (Eq.~\ref{eq:exIntd}), \begin{equation}\label{eq:cnFT2} c_n \left[ \phi (x) \right] = {1 \over Z_0}{\left( -1 \right)^n \over n!} e^{- \int {\rm d^d} x \left[ {1 \over 2} \left( \partial_{\mu} \phi \right)^2 + {1 \over 2} m^2 \phi^2 \right] + n \ln{\left[ \left( \lambda/4 \right) \int {\rm d^d} x \ \phi^4 \right] } } \end{equation} where we have written the $n^{{\rm th}}$ power of the interaction in exponential form. The mathematical analysis below follows closely the discussions in chapter 38 Ref.~\cite{zinn}. Although the problem treated there is different from the one treated here, many techniques used in~\cite{zinn} can be directly borrowed here. For $n$ large enough, the analysis of its ``shape" reduces to the familiar procedure of finding its maxima, as in the case of the simple integrand. The equation determining the maxima of $c_n [ \phi (x) ]$ is the equation that minimizes the exponent, and can be thought of as the equation of motion of the effective action \begin{equation}\label{eq. effactQM} S \left[ \phi \right] = \int {\rm d^d} x \left[ {1 \over 2} \left( \partial_{\mu} \phi \right)^2 + {1 \over 2} m^2 \phi^2 \right] - n \ln{ \left[ {\lambda \over 4} \int {\rm d^d} x \ \phi^4 \right] }, \end{equation} which is \begin{equation}\label{eq:eqQM} - \nabla^2 \phi + m^2 \phi - { 4 n \over \int {\rm d^d} x \phi^4 } \phi^3 = 0. \end{equation} Making the change of variables \begin{equation}\label{eq:chvarQM1} \phi ( x ) = m \left({ \int {\rm d^d} x \phi^4 \over 4n} \right)^{1/2} \varphi ( m x ) = m^{ {\rm d}/2 -1} \left( {4n \over \int {\rm d^d} u \varphi^4 (u) } \right)^{1/2} \varphi ( m x ), \end{equation} $ \varphi $ satisfies the equation \begin{equation}\label{eq:numeqQM} - \nabla^2 \varphi (u) + \varphi (u) - \varphi^3 (u) = 0 \quad , \qquad u \equiv m x. \end{equation} This equation corresponds to the instanton equation of the negative mass $\lambda \phi^4$ theory. The analysis of their solutions can be found in many places. We are interested in the solutions with minimal, finite action. For these solutions, in the infinite volume limit, scaling arguments provide very interesting information. We mentioned at the beginning of section~\ref{sec:wrong} that we work in a finite volume. However, if the volume is large enough, the following infinite volume arguments remain valid up to errors that go to zero exponentially fast when the volume goes to infinity. Since the solution $\phi_{{\rm max}} (x)$ (the subindex ``max" indicates that, in functional space, $c_n \left[ \phi (x) \right]$ reaches its maximum at $\phi_{{\rm max}} (x)$, this should not be confused with the fact that the the action (\ref{eq. effactQM}) reaches its {\it minimum} there) is a minimum of the action (\ref{eq. effactQM}), given an arbitrary constant $\alpha$, $S \left[ \alpha \phi_{{\rm max}} (x) \right]$ should have a minimum at $\alpha = 1$ \cite{zinn,derr}. This implies the equation \begin{equation}\label{Salphphi1} \int {\rm d^d} x \left( \partial_{\mu} \phi_{{\rm max}} \right)^2 + m^2 \int {\rm d^d} x \phi_{{\rm max}}^2 - 4n = 0. \end{equation} Similarly, $S \left[ \phi_{{\rm max}} (\alpha x) \right]$ should also have a minima at $\alpha = 1$, implying, \begin{equation}\label{Salphphi2} { \left( 2 - d \right) \over d} \int {\rm d^d} x \left( \partial_{\mu} \phi_{{\rm max}} \right)^2 - m^2 \int {\rm d^d} x \phi_{{\rm max}}^2 + 2 n = 0. \end{equation} Solving the system of equations (\ref{Salphphi1}) and (\ref{Salphphi2}) we obtain, \begin{eqnarray}\label{solusysteq1} \int {\rm d^d} x \left( \partial_{\mu} \phi_{{\rm max}} \right)^2 &=& n \ d \\ \label{solusysteq2} m^2 \int {\rm d^d} x \phi_{{\rm max}}^2 &=& n \left( 4 - d \right). \end{eqnarray} From which we conclude in particular \begin{equation}\label{eq. effactQMFree} \int {\rm d^d} x \left[ {1 \over 2} \left( \partial_{\mu} \phi_{{\rm max}} \right)^2 + {1 \over 2} m^2 \phi_{{\rm max}}^2 \right] = 2n, \end{equation} independent of the dimension. The relations (\ref{solusysteq1}) and (\ref{solusysteq2}) can be explicitly checked in the case $d=1$ (Quantum Mechanics), in which the solutions to eq.(\ref{eq:eqQM}) are known analytically. They are \begin{equation}\label{d=1sol} \phi_{{\rm max}}^{{\rm d} =1} (t) = \left( {3n \over 2 m} \right)^{1/2} {1 \over \cosh{ \left[ m \left( t - t_0 \right) \right]} } \end{equation} giving \begin{eqnarray}\label{solusysteq1D1} \int {\rm d} t \left( \dot{ \phi}_{{\rm max}}^{{\rm d} =1} \right)^2 &=& n \\ \label{solusysteq2D1} m^2 \int {\rm d} t \left( \phi_{{\rm max}}^{{\rm d} =1} \right)^2 &=& 3 n . \end{eqnarray} Since $\varphi (u)$, introduced in eq.(\ref{eq:chvarQM1}) and satisfying eq.(\ref{eq:numeqQM}), is dimensionless (remember that $u = m x$ is also dimensionless), and the corresponding $\phi_{{\rm max}} (x)$ has finite action, the quantity \begin{equation}\label{Adef} A \equiv { 1 \over 4} \int {\rm d^d} u \varphi^4 (u) \end{equation} is a finite, pure number greater than zero~\cite{zinn}. For the Quantum Mechanical case mentioned above, $A = 4/3$. For the cases $d>1$, $A$ is not explicitly known but, as just said, it must be a finite, positive, pure number. With the definition~(\ref{Adef}), Eq.~(\ref{eq:chvarQM1}) becomes, \begin{equation}\label{relphimaxvarphi} \phi_{{\rm max}} (x) = m^{d/2 - 1} \left( { n \over A} \right)^{1/2} \varphi (m x) \end{equation} Since $\varphi (m x)$ satisfies the n-independent Eq.~(\ref{eq:numeqQM}), we conclude that the field strength of $\phi_{{\rm max}}$ grows with the square root of the order $n$. Equation (\ref{eq. effactQMFree}), together with the definition (\ref{Adef}) and the relation (\ref{relphimaxvarphi}), allow us to write an expression for the action (\ref{eq. effactQM}) at $\phi = \phi_{{\rm max}}$, \begin{equation}\label{Sphimax} S \left[ \phi_{{\rm max}} \right] = 2 n - n \ln{ \left[ {\lambda \ m^{d-4} \over A} n^2 \right] } \end{equation} The value of $c_n \left[ \phi (x) \right]$ at $\phi = \phi_{{\rm max}}$ then becomes, for large n, \begin{equation}\label{cnphimax} c_n \left[ \phi_{{\rm max}} (x) \right] \approx {1 \over Z_0} { \left( -1 \right)^n \over 2 \pi} \left( n - 1 \right) ! \left( {\lambda \ m^{d-4} \over A} \right)^n. \end{equation} With the change of variables \begin{equation}\label{chvarphiq} \phi (x) = \phi_{{\rm max}} (x) + m^{d/2 - 1} \phi_{\rm q} (m x), \end{equation} the Gaussian approximation of $c_n \left[ \phi \right]$ around $\phi_{{\rm max}}$ is, \begin{eqnarray}\label{eq:gaussapp} c_n \left[ \varphi (u) \right] & \approx & {1 \over Z_0} { \left( -1 \right)^n \over 2 \pi} \left( n - 1 \right) ! \left( {\lambda \ m^{d-4} \over A} \right)^n \cdot \nonumber \\ \label{local} & & e^{- {1\over 2} \int {\rm d^d} u_1 {\rm d^d} u_2 \phi_{\rm q} (u_1) \left[ \left( - \nabla^2_{u_1} + 1 - 3 \varphi^2 (u_1) \right) \delta (u_1 - u_2)\right] \phi_{\rm q} (u_2) } \cdot \\ \label{nonlocal} & & e^{- {1\over 2} \int {\rm d^d} u_1 {\rm d^d} u_2 \phi_{\rm q} (u_1) \left[ \left( 1 / A \right) \varphi^3 (u_1) \varphi^3 (u_2) \right] \phi_{\rm q} (u_2) } \end{eqnarray} where $u = mx$ and $\varphi (u)$, solution of Eq.(\ref{eq:numeqQM}), is related to $\phi_{{\rm max}}$ through Eq.(\ref{relphimaxvarphi}). This Gaussian approximation becomes exact in the limit $n \rightarrow \infty$. The second derivative operator, that we call $D$, is then, \begin{equation}\label{D} D = D_{{\rm local}} + D_{{\rm non-local}} \end{equation} with \begin{equation}\label{Dlocal} D_{{\rm local}} = - \nabla^2 + 1 - 3 \varphi^2 \end{equation} and \begin{equation}\label{Dnonlocal} D_{{\rm non-local}} = { 1 \over A} |v> <v| , \quad {\rm with}\quad < u | v > = \varphi^3 (u) \end{equation} and $A$ given in Eq.(\ref{Adef}). The operator $D_{{\rm local}}$ is well known (see for example~\cite{zinn}). It has $d$ eigenvectors $| 0_{\mu} >$ with zero eigenvalues given by \begin{equation}\label{zeroeig} < u | 0_{\mu} > = {\partial \over \partial u^{\mu}}{ \varphi (u) }. \end{equation} These vectors are also zero-eigenvectors of $D$, as can be seen by noting that $|v>$ is orthogonal to them, \begin{equation}\label{orthv0} < v | 0_{\mu} > = 0. \end{equation} They reflect the translation invariance of the action (\ref{eq. effactQM}). $D_{{\rm local}}$ is also known to have one and only one negative eigenvector. The proof of this fact given in Appendix 38 of reference~\cite{zinn}, that uses Sobolev inequalities, can be repeated line by line to prove that, on the contrary, $D$ is a positive semi-definite operator, \begin{equation}\label{Dpossemidef} D \ge 0 \end{equation} in the operator sense. Projecting out the d-dimensional eigenspace of eigenvalue zero, the resulting operator, that we call $D'$, is positive definite. \begin{equation}\label{Dposdef} D' = D'_{{\rm local}} + D_{{\rm non-local}} > 0 \end{equation} This equation explicitly states that the projection over the strictly positive eigenvectors modifies only $D_{{\rm local}}$. The non-local part, as we saw, is a projector orthogonal to the zero modes and is therefore not modified under that operation. Equations (\ref{Dpossemidef}) and (\ref{Dposdef}) suggest that the operator $D$, with the corresponding renormalization for $d>1$, generate a well defined Gaussian measure in a finite volume (remember $d < 4$). In fact, the determinant of $D'_{{\rm local}}$ was calculated many times in the past~\cite{zinn}, and a generalization of a Quantum Mechanical argument of ref.~\cite{au} indicates that this is all we need to compute the determinant of $D'$. The argument goes as follows, \begin{eqnarray}\label{det1} {\rm Det} \left[ D' \right] &=& {\rm Det} \left[ D'_{{\rm local}} + { 1 \over A} |v> <v| \right] \nonumber \\ &=& {\rm Det} \left[ D'_{{\rm local}} \right] \left( 1 + { 1 \over A} <v| D^{' -1}_{{\rm local}} |v> \right). \end{eqnarray} Since $\varphi (u)$ is orthogonal to $\partial_{\mu} \varphi (u)$ (the zero modes of $D$ and $D_{{\rm local}}$), \begin{equation}\label{det2} D'_{{\rm local}} \varphi = D_{{\rm local}} \varphi = -2 \varphi^3. \end{equation} The last equality follows from the definition of $D_{{\rm local}}$ in Eq.(\ref{Dlocal}) and the equation (\ref{eq:numeqQM}) satisfied by $\varphi$. Inverting $D'_{{\rm local}}$, and remembering the definition of $|v>$ and $A$ in Eqs. (\ref{Dnonlocal}) and (\ref{Adef}), we obtain \begin{equation}\label{det3} <v| D^{' -1}_{{\rm local}} |v> = -2 A. \end{equation} Replacing this result in Eq.(\ref{det1}), we arrive at the result \begin{equation}\label{det4} {\rm Det} \left[ D' \right] = - {\rm Det} \left[ D'_{{\rm local}} \right]. \end{equation} As already mentioned, $D'_{{\rm local}}$ has one and only one negative eigenvector, consequently its determinant is negative. Eq. (\ref{det4}) indicates then that ${\rm Det} \left[ D' \right]$ is positive, as it should be according to (\ref{Dposdef}). The effect of the nonlocal part is to change the sign of the determinant of the local part. The preceding equations allow us to integrate the Gaussian approximation of $c_n \left[ \varphi (u) \right]$ given in Eqs.~(\ref{local},\ref{nonlocal}). Using the method of collective coordinates to project out the zero modes, the Jacobian of the corresponding change of variables is, at leading order in $1/ n$, \begin{equation}\label{Jac} J = \prod_{\mu = 1}^{d}{ \left[ \int \left( \partial_{\mu} \phi_{\rm max} \right)^2 {\rm d^d} x \right]^{1/2} } \end{equation} where no sum over $\mu$ is implied. It can be shown that the solutions of Eq.~(\ref{eq:eqQM}) corresponding to minimal action are spherically symmetric~\cite{zinn}, then (\ref{Jac}) can be written as \begin{equation}\label{Jac2} J = \left[ {1 \over {\rm d} } \int \left( \partial_{\mu} \phi_{\rm max} \right)^2 {\rm d^d} x \right]^{{\rm d} /2} \end{equation} where now, sum over $\mu$ is implied. Using Eq.~(\ref{solusysteq1}) we then find \begin{equation}\label{Jac3} J = n^{ {\rm d} /2}. \end{equation} With this expression, the functional integral of $c_n \left[ \varphi (u) \right]$ can be written as \begin{eqnarray}\label{CalcFIcn1} {1 \over Z_0} \int \left[ {\rm d} \phi \right] \ c_n \left[ \phi \right] &=& { \left( -1 \right)^n \over 2 \pi} \left( n - 1 \right) ! \left( {\lambda \ m^{d-4} \over A} \right)^n \cdot \\ \label{CalcFIcn2} & & \left( {\rm Vol} \ m^d \right) n^{d/2} \left( - {\rm Det} \left[ {D'_{{\rm local}} \over D_0} \right] \right)^{-1/2} \end{eqnarray} where$D_0 \equiv - \nabla^2 + 1$. The factors in the line~(\ref{CalcFIcn1}) correspond to the value of $c_n \left[ \phi \right]$ at $\phi_{\rm max}$ up to the normalization $1 / Z_0$ as can be seen in Eq.~(\ref{cnphimax}). The factor ``Vol" arises after the integration over the flat coordinates corresponding to the center of $\phi_{\rm max}$. The $n^{d/2}$ comes from the Jacobian of the change of variables as mentioned before. The factor $m^d$ arises after the rescaling of the fields that makes them dimensionless in both $c_n \left[ \phi \right]$ and $Z_0$. This happens because there are $d$ more integration variables in $Z_0$ due to the integration over the collective coordinated in the numerator. Finally, the factor $\left( - {\rm Det} \left[ D'_{{\rm local}} \right] \right)^{-1/2}$ is the result of the integration over the coordinates orthogonal to the zero modes of $D$, while $\left( {\rm Det} \left[ D_0\right] \right)^{1/2}$ is the dimensionless normalization factor (the mass dimension of both, the numerator and the denominator, was already taking care of in the term $m^d$). The minus sign is due to the non-local part of $D$ that, as proved above, simply changes the sign of the determinant of the local part, making it positive. In the case $d=1$, $- {\rm Det} \left[ D'_{{\rm local}} / D_0 \right] = 1/ 12$~\cite{zinn,au}, and Eqs.~(\ref{CalcFIcn1},\ref{CalcFIcn2}) (with $A= 4/3 $ as already mentioned) become identical to the corresponding result of Ref.~\cite{au} if we take into account the different normalization here and a factor of 2 that is taken care of by remembering that the sign of the solution of Eq.~(\ref{eq:eqQM}) is undetermined, therefore both, positive and negative solutions contribute equally to the functional integral. For $d = 2$ or 3, the formal expression~(\ref{CalcFIcn1},\ref{CalcFIcn2}) needs of course to be renormalized. All the arguments in this section remain valid for the theory with a Pauli-Villars regularization~\cite{zinn}. The action~(\ref{eq:action}) becomes \begin{equation} S \left[ \phi \right] = \int {\rm d^d} x \left[ {1 \over 2} \phi \left( - \nabla^2 + {\nabla^4 \over \Lambda^2} + m^2 \right) \phi + {\lambda \over 4} \phi^4 + {1 \over 2} \delta m^2 (\Lambda) \ \phi^2 \right]. \label{eq:actionregul} \end{equation} The modification of the kinetic part of the action affects both the equation~(\ref{eq:eqQM}) and the scaling arguments, but by an amount that becomes small like $\Lambda^{-2}$ when the ultra-violet cut-off $\Lambda$ becomes large. As shown in Ref.~\cite{zinn}, although the counterterm increases with the cut-off, since it is proportional to at least one power of $\lambda$, taking the small $\lambda$ limit before the large cut-off limit justifies to ignore it in the equation~(\ref{eq:eqQM}) and the scaling arguments. On the other hand it contributes to the result~(\ref{CalcFIcn1},\ref{CalcFIcn2}) an amount that exactly cancels the divergence in the ${\rm Det} \left[ D'_{{\rm local}} \right]$ making the final expression finite as it should be. In the large $n$ limit, where the Gaussian approximation~(\ref{local},\ref{nonlocal}) becomes exact, the expression~(\ref{CalcFIcn1},\ref{CalcFIcn2}) gives the large order behavior of the perturbative series of $Z$ (up to the factor of 2 mentioned above) {\it without any assumption about the analytic structure in $\lambda$}~\cite{au}. A completely analogous procedure would give the large order behavior of any Green's function. Eqs.~(\ref{relphimaxvarphi}),~(\ref{cnphimax}),~(\ref{local},\ref{nonlocal}) and~(\ref{CalcFIcn1},\ref{CalcFIcn2}) , allow us to draw an accurate picture of the mechanism underlying the lack of domination (in the sense of the Lebesgue's theorem) of the convergence of the sequence of perturbative integrands~(\ref{eq:fNfunctint}) towards~(\ref{eq:exIntd}), and consequently of the mechanism underlying the divergence of the perturbative series. In fact, perhaps not surprisingly given the similarity of their large order behavior, this picture is very similar to the one described in the previous section for the simple integral example. In a finite volume, there is a region of finite measure in field space in which the perturbative approximation $f_N [ \phi (x) ]$ of Eq.~(\ref{eq:fNfunctint}) approximate the exact integrand~(\ref{eq:exIntd}) with an error smaller than a given prescribed number. This region grows with $N$, becoming the full field space in the $N \rightarrow \infty$ limit. As in the simple example, this is a consequence of Taylor's theorem applied to the (analytic) integrand~(\ref{eq:exIntd}). The problem is that, for any finite $N$, outside that region the approximate integrand $f_N [ \phi (x) ]$ strongly deviates from the exact one. This can be seen by noting that the maxima of every term of $f_N$ grow factorially with the order. Therefore, for large enough $N$, the last term is far greater than the previous ones at its maxima. Even more, as shown above, the Gaussian approximation around that maxima (that becomes exact for $N \rightarrow \infty$) defines a measure that does not go to zero as $N \rightarrow \infty$ (in fact, it is independent of $N$~(\ref{local},\ref{nonlocal})). This means that for every finite $N$, there is a region of finite measure in field space (and this measure does not go to zero as $N \rightarrow \infty$) in which the deviation between the perturbative integrand and the exact one is of the order of the maxima of the last term of $f_N$, i.e., of the order of $\left( N - 1 \right)!$. No integrable functional can therefore satisfy the property (\ref{eq:bound}) of the Lebesgue's theorem. That is the mechanism that makes the sequence of perturbative integrands, although convergent to the exact one, non-dominated in the sense of Lebesgue's theorem. That is therefore the mechanism that makes the sequence of integrals (i.e., the perturbative series) divergent. In fact, as Eqs.~(\ref{CalcFIcn1},\ref{CalcFIcn2}) show, the famous factorial behavior of the large order coefficients of the perturbative series is a consequence, after integration, of the above mechanism. \section{Steps Towards a Convergent Series}\label{sec:converg} It was mentioned in the introduction that the analysis of the divergence of perturbation theory presented in this paper would point directly towards the aspects of the perturbative series that need to be modified in order to generate a convergent series. This is the topic of the present section. In the previous section we analyzed perturbation theory from the point of view of the Dominated Convergence Theorem. We have detected the precise way in which the convergence of the sequence of perturbative integrands to the exact one takes place, and the way this convergence fails to be dominated. We have learned that for any finite order $N$, the field space naturally divides into two regions. In the first one, that grows with the order, eventually becoming the full field space (in the $N \rightarrow \infty$ limit), the perturbative integrands very accurately approximate the exact one. In the other one, however, the deviation between the perturbative and exact integrands is so strong, that the sequence of integrals diverge. It is then clear that {\it if we could somehow modify the integrands, order by order, in the region where they deviate from the exact one, while preserving them as they are in the other region, then, with a ``proper" modification, such modified sequence of integrands would converge in a dominated way. According to the Dominated Convergence Theorem, their integrals would then converge to the exact integral, achieving the desired goal of a convergent modified perturbation theory.} \\ Let us call $\Omega_N$ to the region of field space in which the ${\rm N^{th}}$ perturbative integrand approximate with a given prescribed error the exact integrand (\ref{eq:exIntd}). The {\it characteristic function}, ${\rm Ch} (\Omega_N, \left\{ \phi (x) \right\})$, of that region is equal to 1 for field configurations belonging to it, and zero otherwise: \begin{equation}\label{chfunctome} {\rm Ch} (\Omega_N, \left\{ \phi (x) \right\}) \equiv \cases{ 1 & for $ \left\{ \phi (x) \right\} \in \Omega_N $ \cr 0 & for $ \left\{ \phi (x) \right\} \not\in \Omega_N $} \end{equation} One possible realization of the above strategy of modifying the integrands (\ref{eq:fNfunctint}) in the ``bad" region of field space is to make them zero there. We would have \begin{equation}\label{fprimach} f_N' \left[ \phi (x) \right]= {1 \over Z_0} \sum_{n=0}^{N} {\left( -1 \right)^n \over n!} e^{- S_0 } \left( {\lambda \over 4} \int {\rm d^d} x \phi^4 \right)^n {\rm Ch} (\Omega_N, \left\{ \phi (x) \right\}) \end{equation} According to the analysis of the previous section, choosing $\Omega_N$ appropriately, the sequence of $f_N' \left[ \phi (x) \right]$ will converge dominatedly, and the corresponding interchange between sum and integral will now be allowed. A rigorous proof of this is left for a paper currently in preparation. For the purposes of the present argument, it is sufficient to rely on the analysis of the previous section to assume its validity. Also, in the next section we will analyze, along the general ideas of this paper, some resummation schemes for which rigorous proofs of convergence have recently been given~[12-19]. As that analysis will show, these methods strongly rely on the general notions underlying the formula~(\ref{fprimach}). Their convergence supports, then, the validity of the dominated nature of the convergence of~(\ref{fprimach}) towards~(\ref{eq:exIntd}). An urgent issue, however, is the practical applicability of the above strategy. To implement it, we need a functional representation of the characteristic function~(\ref{chfunctome}) (or an approximation to it) that only involves {\it Gaussian and polynomial} functionals. In the same way in which a functional representation of the Dirac delta function allow us to perform functional integrals with constraints, the Fadeev-Popov quantization of Gauge theories being the most famous example, a functional representation of the characteristic function~(\ref{chfunctome}) would allow us to functionally integrate only the desired region of functional space. Since, basically, the functionals we know how to integrate reduce to Gaussians multiplied by polynomials, the desired representation of the characteristic function should {\it only} involve those functionals. Conversely, if it only involves those functionals, all the sophisticated machinery developed for perturbation theory (including all the perturbative renormalization methods) would automatically be applicable. With this in mind, consider the following function, \begin{equation}\label{win} W (M, u) \equiv e^{-Mu} \sum_{j=0}^{M} {\left( Mu \right)^j \over j! } \end{equation} where $M$ is a positive integer. Note that $W (M,u)$ arises from $1 = e^{-Mu} e^{+Mu}$ by expanding the second exponential up to order M. $W (M, u)$ has the following remarkable properties, \begin{enumerate} \item $W(M,u) \rightarrow 1 $ when $M \rightarrow \infty$ for $0<u < 1$. The convergence is uniform, with the error going to zero as \begin{equation}\label{item1} R (M,u) \le e^{M \left( \ln{u} - (u - 1) \right)} {1 \over \sqrt{2 \pi M}} {u \over 1 - u + 1/M}. \end{equation} \item $W(M,u) \rightarrow 0 $ when $M \rightarrow \infty$ for $1<u $. The convergence is also uniform , with an error of the form \begin{equation}\label{item2} W (M, u) \le e^{M \left( \ln{u} - (u - 1) \right)}. \end{equation} \end{enumerate} As we see, the exponent corresponds to the same function in both cases. For $u > 0$, this function is always negative except at its maxima, at $u =1$, where it is 0. Therefore the convergence is in both cases exponentially fast in $M$, with the exponent becoming more and more negative, for a fixed $M$, when $u$ differs more and more from 1. The proof of properties 1 and 2 is in the Appendix 1. If we replace $u$ by a positive definite quadratic form $< \phi | D | \phi >/C_N$, then the insertion of Eq.~(\ref{win}) into the functional integral would effectively cut off the region of integration $< \phi | D | \phi > \ > C_N$ \begin{eqnarray}\label{fprimaw} Z_N' \left[ \phi (x) \right] &=& {1 \over Z_0} \int \left[ {\rm d} \phi \right] \sum_{n=0}^{N} {\left( -S_{\rm Int} \right)^n \over n!} e^{- S_0 } \lim_{M \rightarrow \infty} W \left( M, {< \phi | D | \phi> \over C_N} \right) \\ &=& {1 \over Z_0} \sum_{n=0}^{N} {\left( -1 \right)^n \over n!} \lim_{M \rightarrow \infty} \int \left[ {\rm d} \phi \right] e^{- S_0 } \left( S_{\rm Int} \right)^n W \left( M, {< \phi | D | \phi> \over C_N} \right) \nonumber \\ \label{fprimaw2} \end{eqnarray} $C_N$ is a constant that changes with the order $N$ of the expansion in $\lambda$, increasing with $N$ but in such a way that in the region $< \phi | D | \phi> < C_N$ the difference between the perturbative and the exact integrands is smaller than a given prescribed error. Since the convergence of $W$ is uniform according to properties 1 and 2, with errors given in Eqs.~(\ref{item1}) and~(\ref{item2}), the corresponding interchange between the sum in Eq.~(\ref{fprimaw2}) and the functional integral is justified. The fact that $u$ becomes a {\it quadratic} form implies that the resulting integrands are Gaussians multiplied by monomials, therefore the familiar Feynman diagram techniques can be used to integrate them. It also implies that no new loops appear and the sum in $j$ from~(\ref{win}) becomes an algebraic problem. A typical functional integral to compute has the form \begin{equation}\label{typicalfunctint} \int \left[ {\rm d} \phi \right] e^{- \int {\rm d^d} x \left[ {1 \over 2} \left( \partial_{\mu} \phi \right)^2 + {1 \over 2} m^2 \phi^2 + ( \phi D \phi / C_N ) \right] } \left( \int {\rm d^d} x \phi^4 \right)^n \left( \int {\rm d^d} x \ \phi D \phi \right)^m \end{equation} as can be seen by replacing the definition~(\ref{win}) into~(\ref{fprimaw2}) with $u = < \phi | D | \phi > / C_N$. Note that at any given order in $\lambda$, it is not necessary in principle to go to infinity in $M$. That would amount to replace the perturbative integrands by zero in the region $< \phi | D | \phi > \ > C_N$, realizing the strategy mentioned before. But since the convergence in $W$ is uniform, a finite, large enough $M$ (depending on the order in the expansion in the coupling constant), would suffice to tame the behavior of the perturbative integrands and transform them into a {\it dominated} convergent sequence. In fact, as we will see, many methods of improvement of perturbation theory use effectively formula~(\ref{win}) without sending $M \rightarrow \infty$ for any given finite order in perturbation theory. In any case, as already mentioned, that limit is in principle computable, since it does not involves new loops. Work in this direction is in progress. The convergence of the sequence~(\ref{fprimaw2}) towards $Z (\lambda)$ may be thought, at first sight, to be in conflict with our well established knowledge about the non-analyticity of this function at $\lambda = 0$. In fact, Eq.~(\ref{fprimaw2}) seems to be a power series in $\lambda$ (the powers of $\lambda$ coming from the powers of $S_{\rm Int}$), therefore, if convergent, that power series would define a function of $\lambda$ analytic at $\lambda = 0$. It must be recognized, however, that the validity of Lebesgue's Dominated Convergence Theorem is completely independent of any analyticity consideration. Therefore, if its hypothesis are satisfied, its conclusions must be valid. This being said, the question of how does the convergence of~(\ref{fprimaw2}) fits with the non-anlayticity of $Z ( \lambda )$ deserves an answer. To begin with, even at finite order in $\lambda$, the function~(\ref{fprimaw2}) is not necessarily analytic at $\lambda =0$ despite its analytic appearance. This is because the constant $C_N$ may have an implicit nonanalytic dependence on $\lambda$. In Appendix 2 this is actually the case in the context of a simple example to which the present ideas are applied. But the mechanism that ultimately introduces the proper non-analyticity in $\lambda$ is the limit process $N \rightarrow \infty$. Given a non-analytic function like $Z (\lambda)$ one can always construct a sequence of analytic functions that converge to it. Satisfying the hypothesis of the Dominated Convergence Theorem is a way of achieving that, avoiding all the complicated and {\it model dependent} issues of non-analyticity. Note that the validity of these hypothesis for a given sequence of integrands can be checked independently of any analyticity consideration. In the Appendix 2 we prove the convergence of the general strategy discussed here for the simple integral example analyzed in section~(\ref{sec:simpleint}). For that case, making $u = (x/x_{c,N})^2$, the function $W (M, u)$ becomes in the limit the characteristic function of the interval $|x|< x_{c,N}$. We use this to explicitly compute the non-analytic function $z (\lambda)$ (Eq.~(\ref{eq:pertsimple})) calculating only Gaussian integrals. We also show explicitly how a non-analytic dependence of $x_{c,N}$ on $\lambda$ naturally arises just by demanding the validity of the Lebesgue's hypothesis and how the $N \rightarrow \infty$ limit process captures the full non-analyticity of $z (\lambda)$. The same method also works for the ``negative mass case", where the Borel resummation method fails. In figure~\ref{fig3}, we can appreciate the convergence of $W$ towards the characteristic function of the interval $|x|< x_{c,N}$ for $x_{c,N} = 1$ for two different values of $M$. \begin{figure}[htp] \hbox to \hsize{\hss\psfig{figure=win.eps,width=0.7\hsize}\hss} \caption{Function $W(M,x,x_{c,N})$ with $x_{c,N}=1$ for $M=3$ (segmented line) and $M=60$ (continuous line). The convergence towards the characteristic function of the interval $|x|< x_{c,N}$ is apparent.} \label{fig3} \end{figure} \section{Improvement methods of perturbation theory.}\label{Improvement} The analysis of the mechanism of divergence of the perturbative series presented in this paper, together with the formula~(\ref{win}) and its properties, offer a large range of possibilities to construct a convergent series. In the previous section we have shown how that formula can be used to effectively cut-off the region of field space where the strong deviation between perturbative and exact integrands take place. But as we will see, this is only one possible way, among many, to use the formula~(\ref{win}) to transform the sequence of perturbative integrands into a dominated one. Another example of its possible use is the so called ``optimized delta expansion"~\cite{gral, stev}. In a series of papers~\cite{tony1}-\cite{kle}, it was proved that such an expansion converges for the partition function of the anharmonic oscillator in finite Euclidean time. The problem of convergence in the infinite Euclidean time (or zero temperature) limit for the free energy or any connected Green's function is still under investigation, as well as its extension to Quantum Field Theories~\cite{tony3,arv}. The method was proved to generate a convergent series for the energy eigenvalues~\cite{gui1, gui2}, although such studies make heavy use of analyticity properties valid specifically in the models studied. In these works, it was realized that many methods of improvement of perturbations theory, such as the order dependent mappings of references~\cite{zinn2, leg}, posses the same general structure as the linear delta expansion. A considerable amount of work has been dedicated to investigate the virtues and limitations of the method and extensions of it~\cite{kle, nev}. It is not the place here to give a detailed analysis of these methods. But we would like to briefly indicate how they can be understood in terms of the ideas presented here. In what follows, our analysis is restricted to $d=1$ (Quantum Mechanics) where rigorous results about the convergence of the methods considered here are available. Let's consider the case of the anharmonic oscillator. Its action is given in Eq.~(\ref{eq:action}) for $d=1$. The idea of the method is to replace it by an interpolating action \begin{equation}\label{sdelta} S_{\delta} = \int {\rm d} t \left[ {1 \over 2} \left( {\rm d}_t \phi \right)^2 + {1 \over 2} \left( m^2 + {\lambda \over 2 m} \alpha \right) \phi^2 + \delta {\lambda \over 4} \left( \phi^4 -{ \alpha \over m} \phi^2 \right) \right]. \end{equation} Clearly, the dependence on the parameter $\alpha$ in $S_{\delta}$ is lost when $\delta = 1$. For that value, the action~(\ref{sdelta}) reduces to~(\ref{eq:action}). However, if we expand up to a finite order in $\delta$ and then make $\delta=1$, the result still depends on $\alpha$. The idea is to tune $\alpha$, order by order in the expansion in $\delta$, so that the result is a convergent series. It was shown in the references mentioned above that the methods works if $\alpha$ is tuned properly. For example, in reference~\cite{tony2}, the asymptotic scaling $\alpha \simeq N^{2/3}$ was used to prove the convergence of the method for the partition function at finite Euclidean time. It is interesting to note that originally~\cite{tony1, tony2}, $\alpha$ was tuned according to heuristic prescriptions such as the ``principle of minimal sensitivity"~\cite{stev} (at any given order in $\delta$, choose $\alpha$ so that the result is insensitive to small changes in it), or the criterion of ``fastest apparent convergence" (the value of $\alpha$ at which the next order in delta vanishes). But later~\cite{tony3, arv}, it was realized that the best strategy was simply to leave $\alpha$ undetermined, find an expression for the error (that obviously depends on $\alpha$), and then choose $\alpha$ so that the error goes to zero when the order in $\delta$ goes to infinity. It is clear that a {\it structural} understanding of the convergence of the method can help to construct the necessary generalizations to overcome the difficulties associated with the convergence in the infinite volume limit for connected Green's functions, as well as the extensions to general Quantum Field Theories. To understand the ``optimized delta expansion" in terms of the ideas presented in this paper, let us expand the functional integral corresponding to the action~(\ref{sdelta}) in powers of $\delta$ up to a finite order $N$, and make $\delta =1$ as the method indicates, \begin{eqnarray}\label{delta1} Z (m, \lambda, \alpha, N) &=& {1 \over Z_0} \int \left[ {\rm d} \phi \right] e^{- \int {\rm d} t \left[ {1 \over 2} \left( {\rm d}_t \phi \right)^2 + {1 \over 2} \left( m^2 + {\lambda \alpha \over 2 m} \right) \phi^2 \right] } \cdot \nonumber \\ & & \left[ \sum_{n=0}^N {\left( - 1 \right)^n \over n!} \left( {\lambda \over 4} \int \phi^4 - {\lambda \alpha \over 4 m} \int \phi^2 \right)^n \right] \end{eqnarray} The general analysis of the mechanism of divergence of perturbation theory of section~\ref{sec:leb} indicates that if the function~(\ref{delta1}) generates a convergent series with $\alpha$ scaling properly with $N$, then, barring miraculous coincidences, the corresponding integrands should converge dominatedly (or, even better, uniformly) towards the exact integrand~(\ref{eq:fNfunctint}). We want to obtain a qualitative understanding on how this method achieves that. Expanding the binomial and making some elementary changes of variables in the indices of summation, we obtain the expression \begin{eqnarray}\label{delta2} Z (m, \lambda, \alpha, N) &=& {1 \over Z_0} \int \left[ {\rm d} \phi \right] e^{- \int {\rm d} t \left[ {1 \over 2} \left( {\rm d}_t \phi \right)^2 + {1 \over 2} \left( m^2 + {\lambda \alpha \over 2 m} \right) \phi^2 \right] } \cdot \nonumber \\ & & \left[ \sum_{i=0}^N {\left( - 1 \right)^i \over i!} \left( {\lambda \over 4} \int \phi^4 \right)^i \left( \sum_{k=0}^{N-i} {1 \over k!} \left( {\lambda \alpha \over 4 m} \int \phi^2 \right)^k \right) \right] \end{eqnarray} This equation already shows some of the distinctive characteristics of the method. As we see, the ${\rm i^{th}}$ power of the interacting action in the expansion of $e^{-S_{\rm Int}}$ up to order $N$, is multiplied by \begin{equation}\label{delta3} {\cal W} \left( N - i \right) \equiv e^{- \left( \lambda \alpha / 4 m \right) \int \phi^2} \left( \sum_{k=0}^{N-i} {1 \over k!} \left( {\lambda \alpha \over 4 m} \int \phi^2 \right)^k \right). \end{equation} Note that ${\cal W} \left( N \right)$ corresponds to the function $W (M, u)$ with $M = N$ ($N$ is the order in the expansion of $e^{-S_{\rm Int}}$), and the variable $u$ replaced by the quadratic form $\left( \left( \lambda / 4 m \right) \int \phi^2 \right) / C_N$, where $C_N = N / \alpha$. Making for example $\alpha \simeq N^{2/3}$ as in Ref.~\cite{tony2} (where it was proved that with such scaling the method generates a convergent series), we see then, that, according to the previous section, ${\cal W} \left( N \right)$ is an approximation of the theta function in the region of field space characterized by \begin{equation}\label{delta4} {\lambda \alpha \over 4 m} \int {\rm d} x \ \phi^2 \le N^{1/3}. \end{equation} Equation~(\ref{delta2}), however, shows that the mechanism used to achieve dominated convergence can not be reduced to a simple insertion of the function $W(M,u)$ with $M=N$ and $u = \left( \left( \lambda / 4 m \right) \int \phi^2 \right) / C_N$. That would be the case if all the powers of the expansion of $e^{-S_{\rm Int}}$ up to order $N$ were multiplied by ${\cal W} \left( N \right)$. But equation~(\ref{delta2}) shows that the ${\rm i^{th}}$ power of the interacting action is in fact multiplied by ${\cal W} \left( N - i \right)$. \\ At this point it is convenient to pause for a moment in our study of the ``optimized delta expansion" to give some useful definitions. Let us call {\bf passive} mechanisms (to achieve dominated, or uniform convergence of a sequence of integrands to the exact one) to those that can be reduced to the product of the $N^{\rm th}$ perturbative integrand and the characteristic function of a region $\Omega_N$ of field space for some sequence $\left\{ \Omega_N \right\}$. Passive methods use only information that is already available in the perturbative integrands, they just get rid of the ``noise" inherent to perturbation theory. Because of that, in addition to define a convergent series, they can also be very useful to study perturbation theory itself. The function $W (N,u)$, with $u$ replaced by a properly selected quadratic operator, was specially designed to make passive methods practical. In a sense, section~\ref{sec:converg} is a discussion of passive methods. {\bf Active} mechanisms are those that are not passive, as defined above. \\ What kind of mechanism is the one underlying the ``optimized delta expansion" method? A trivial generalization of the proof, in the previous section, of the convergence of $W(M,u)$ towards the theta function for $u>0$, shows that the function \begin{equation}\label{wbarra} \overline{W} (M,u,i) \equiv e^{-Mu} \sum_{n=0}^{M-i} {\left( M u \right)^i \over i!} \end{equation} also converges towards the theta function for $u>0$ in the limit \begin{equation}\label{limit} M \rightarrow \infty, \quad i \ {\rm fixed}. \end{equation} In this sense, the ``optimized delta expansion" method does have passive aspects. As Eqs.~(\ref{delta2}, \ref{delta3}) show, it amounts to multiplying the ${\rm i^{th}}$ power of the expansion up to order $N$ of $e^{S_{\rm Int}}$ by $\overline{W}$ with $u = \left( \left( \lambda / 4 m \right) \int \phi^2 \right) / C_N$ and $C_N = N / \alpha$. Since this function converges to the characteristic function of the region characterized by Eq.~(\ref{delta4}), this means that the first $i$ terms, of the expansion up to order $N$ of $e^{S_{\rm Int}}$ are {\it effectively} multiplied by the same function (an approximate characteristic function) for $i \ll N$ . Therefore, the first $i$ terms, with $i \ll N$, use only the information available in the perturbative series to converge to the exact integrand. What about the other terms?, the ones characterized by $i {\ \lower-1.2pt\vbox{\hbox{\rlap{$<$}\lower5pt\vbox{\hbox{$\sim$}}}}\ } N$? Surprisingly, these terms produce a convergence of the corresponding integrands towards the exact one that is faster than possible with only passive components! It is not the place here to study this aspect in detail, so, let us simply see this ``faster than passive" convergence for the simple integral example. Applied to the ``mass-less" version of the integral~(\ref{eq:pertsimple}), the optimized delta expansion method was proved to generate a rapidly convergent sequence in Ref.~\cite{tony1}. That is, the sequence given by \begin{equation}\label{simODE} I_N \equiv \sum_{n=0}^N {(-1)^n \over n!} \int_{- \infty}^{\infty} {\rm d} x \ e^{ - \alpha (N) \ x^2} ({\lambda \over 4} x^4- \alpha (N) \ x^2)^n \end{equation} was proved to converge to \begin{equation}\label{sim} I \equiv \int_{- \infty}^{\infty} {\rm d} x \ e^{ - \lambda x^4 / 4} \end{equation} when $\alpha (N) \simeq \sqrt{N}$ with an error that goes to zero at the very fast rate of $R_N < C N^{1/4} e^{-0.663 N}$ when $N \rightarrow \infty$. $C$ is a numerical constant. We are interested in understanding whether the corresponding convergence of the integrands is faster than passive. For our qualitative purposes, it is enough to observe, in figure~\ref{fig4}, the convergence towards the exact integrand \begin{equation}\label{exinm0} I_{\rm exa} (x) = e^{ - {\lambda \over 4} x^4} \end{equation} of both, the perturbative integrand \begin{equation}\label{pertintm0} I_{\rm pert} = \sum_{n=0}^N { ( - \lambda x^4 / 4)^n \over n!} \end{equation} and the optimized delta expansion integrand \begin{equation}\label{odeinm0} I_{\rm ode} = \sum_{n=0}^N {(-1)^n \over n!} e^{ - \alpha (N) \ x^2} ({\lambda \over 4} x^4- \alpha (N) \ x^2)^n \end{equation} with $\alpha (N) \simeq \sqrt{N}$, for $N=4$. \begin{figure}[htp] \hbox to \hsize{\hss\psfig{figure=odelexp.eps,width=0.9\hsize}\hss} \caption{In the main plot, the superiority of the convergence of the fourth order optimized delta expansion (ode) with respect to the same order perturbative approximation is evident. In the sub-graph, the difference between the ode and the exact integrand is plotted. Note the difference in the scales of the y axis of the main and sub graph.} \label{fig4} \end{figure} We can see how accurate the convergence of $I_{\rm ode} (x)$ is, already at this low order. In particular, when the perturbative integrand begins to diverge, $I_{\rm ode} (x)$ continues to approximate remarkably well the exact integrand. In the sub-figure, we can appreciate the difference between $I_{\rm exa} (x)$ and $I_{\rm ode} (x)$. Note the difference in the $y$ axis scale of graph and sub-graph. It is then clear that the optimized delta expansion method, with its subtle combination of passive and active components, manages to generate a sequence of integrands that (uniformly) converges towards the exact one at a rate that far exceeds the possibilities within a purely passive method. We stop here the qualitative discussion of the optimized delta expansion method expressing two general lessons: \begin{enumerate} \item Any method of improvement of the perturbative series in a given quantum theory, where a functional integral representation of the quantity under study exists, must rely, at the level of the integrands, on an improvement over the pointwise convergence of the Taylor series in the coupling constants of $e^{-S}$. \item The problem of finding a convergent series reduces to the problem of finding a dominatedly convergent sequence of integrands towards $e^{-S}$. \end{enumerate} This second simple statement, not only provides a guide to the construction of convergent schemes, but also emphasizes the fact that, in principle, a dominatedly convergent sequence of integrands, do not have to have any relation whatsoever with the corresponding Taylor expansion. In order to be able to use the usual techniques of Quantum Field Theory, it is reasonable to restrict the search for a convergent scheme to a sequence of integrands of the general form: \begin{equation}\label{generic} f_N = e^{-S_0} \sum_{n=0}^N a_n \ S_{\rm Int}^n \ f_n (< \phi | D_n | \phi >) \end{equation} where the functional $f_n$ of the quadratic form $< \phi | D_n | \phi >$, should take care of the non-dominated convergence that is bound to appear with only powers of the interacting action. The function $W$ of section~\ref{sec:converg}, with its possible generalizations, is an ideal candidate for this purpose. But the selection of the coefficients $a_n$ amounts to a pure problem in optimization of the convergence of the integrands $-$ no a priori connection with any Taylor series is necessary. \section{Conclusions}\label{sec:concl} In this paper we have exposed the mechanism, at the level of the integrands, that make the perturbative expansion of a functional integral divergent. We have seen in detail how the sequence of integrands violate the domination hypothesis of the Lebesgue's Dominated Convergent Theorem. That theorem, as is well known, establishes the conditions under which it is allowed to interchange an integration and a limit, in particular the one that takes place in the generation of perturbation series. It was shown that at any finite order in perturbation theory, the field space divides into two regions. One, that grows with the order, in which the perturbative integrands very accurately approximate the exact integrand. In the other however, a strong deviation takes place. It was shown that the behavior in this second region violates the hypothesis of Lebesgue's theorem, and, consequently, generates the divergence of perturbation theory. The famous factorial growth of the large order coefficients of the perturbative series was shown to be an effect, after integration, of the very mechanism that violates the hypothesis of the theorem. All of the above was done explicitly without relying in the particular analytic properties of the models studied. It is therefore natural to assume that similar mechanisms of violation of the Lebesgue's hypothesis are present in any other Quantum Field Theory, although for just renormalizable theories other mechanisms are responsible for renormalons. Studies in this direction are in progress. The mechanism of divergence presented here points towards a simple way to achieve a convergent series: integrate only in the ``good" region of field space. Since this region grows with the order, becoming in the limit the whole field space, integrating in a correspondingly increasing region we would obtain a convergent series. A step forward towards a practical implementation of this program was made with the construction of the function $W$~(\ref{win}). This function allows us to introduce a Gaussian representation of the characteristic function of regions of field space, very much like the imposition of constraints in the functional integral was allowed by a functional representation of the Dirac's delta function. A rigorous proof of the convergence of this practical implementation of the above mentioned strategy is in progress. In Appendix 2 it was applied to a simple integral example. Finally, a qualitative analysis of the optimized delta expansion method of improvement of perturbation theory in terms of the ideas of this paper was done. Some general properties of improvement methods, useful to generates new schemes, as well as to understand and improve old ones, have been established. \newpage \section*{Acknowledgments} S. P. was supported in part by U.S. Dept. of Energy Grant DE-FG 02-91ER40685. He would like to thank A. Duncan for numerous very useful discussions, and also to A. Das and S. Rajeev. \section*{Appendix 1} In this appendix we will prove the two properties of formula~(\ref{win}). Because for $u>0$ all the terms of the sum defining $W (M, u)$ are positive, we have trivially $W (M, u) > 0$. On the other hand, since in the Taylor expansion of $e^{Mu}$ all the terms are positive, we have $\sum_{n=0}^{M} \left( Mu \right)^n / n! \le e^{Mu}$. Therefore $W (M, u) \le 1$. So for every $M$ and positive or zero $u$ we have, \begin{equation}\label{app:bound} 0 \le W(M,u) \le 1. \end{equation} Consider first the case $0<u < 1$. \begin{equation}\label{app:rest} 1 - W (M, u) = e^{-Mu} \sum_{n=M+1}^{\infty} {\left( Mu \right)^n \over n! } \equiv R(M,u), \end{equation} we will prove that $R(M,u) \rightarrow 0$ when $M \rightarrow \infty$. Changing variables to $j = n -M$, we get \begin{eqnarray}\label{app:rest2} R(M,u) &=& e^{-Mu} {\left( Mu \right)^M \over M! } \sum_{j=1}^{\infty} \left( Mu \right)^j {M! \over \left( j+M \right)! } \\ &\le& e^{-Mu} {\left( Mu \right)^M \over M! } \sum_{j=1}^{\infty} {\left( Mu \right)^j \over \left( M + 1 \right)^j } \\ &\le& e^{-Mu} {\left( Mu \right)^M \over M! } {u \over 1 - u + 1/M}. \end{eqnarray} But $M^M / M! \rightarrow e^M / \sqrt{2 \pi M}$ for large $M$, so \begin{equation}\label{app:rest3} R (M,u) \le e^{M \left( \ln{u} - (u - 1) \right)} {1 \over \sqrt{2 \pi M}} {u \over 1 - u + 1/M}. \end{equation} The exponent is negative in the region $0<u < 1$ since, being both, $\ln{u}$ and $(u - 1)$ negative there, $| \ln{u} | > |u - 1|$ in this region. Therefore \begin{equation}\label{app:proved} R (M,u) \rightarrow 0 \ , {\rm when} \quad M \rightarrow \infty \end{equation} in the region $0<u < 1$ and property 1 is proved with an exponentially fast convergence. In the region $u > 1$, we have \begin{eqnarray}\label{app:uge1} W (M, u) &=& e^{-Mu} \sum_{n=0}^{M} {\left( Mu \right)^n \over n! } \le e^{-Mu} u^M \sum_{n=0}^{M} {M^n \over n! } \\ \label{app:uge2} &\le& e^{M \left( \ln{u} - (u - 1) \right)}. \end{eqnarray} The first inequality is valid because $u>1$ and the second because \\ $e^M > \sum_{n=0}^{M} M^n / n! $. The exponent is again negative. For $u>1$, both, $ \ln{u}$ and $( u-1)$ are positive, but now $| \ln{u} | < |u - 1|$. So property 2 is also valid with an exponentially fast convergence. For $u=1$ all we know is that $W$ is bounded by Eq.(\ref{app:bound}). That is all we need. Numerics suggest $W (M, 1) \rightarrow 1/2$ when $M \rightarrow \infty$. This finishes our proof. \section*{Appendix 2} In this appendix we apply the strategy discussed in section~\ref{sec:converg} to generate a series convergent to the function $z (\lambda)$ (Eq.~(\ref{eq:pertsimple})). This is done using the function $W$ of Eq.~(\ref{win}) and computing {\it exclusively} Gaussian integrals, therefore we restrict ourselves to using only those techniques that are also available in Quantum Field Theory. As mentioned in section~\ref{sec:converg}, the simplest possible modification of the perturbative integrand (\ref{eq:fNsimpltint}) that would transform the corresponding sequence into a dominated one, amounts to keep them as they are for $|x| < x_{c,N}$ and replacing them by zero for $|x| > x_{c,N}$. That is, \begin{equation}\label{convseq} f'_N = \cases{ \pi^{-1/2} \sum_{n=0}^{N} {\left( -1 \right)^n \over n! } \left( {\lambda \over 4} x^4 \right)^n e^{- x^2 } & for $|x| < x_{c,N} $ \cr 0 & for $|x| > x_{c,N}$}. \end{equation} In fact, choosing $x_{c,N}$ so as to properly avoid the region in which the deviation takes place, the sequence of $f'_N$ converges {\it uniformly} towards the exact integrand (\ref{eq:exintd}) as we will show shortly. Consequently, the corresponding sequence of integrals \begin{eqnarray}\label{exint2} \int_{- \infty}^{\infty} {\rm d} x f'_N & = & \pi^{-1/2} \int_{- x_{c,N} }^{x_{c,N}} {\rm d} x \sum_{n=0}^{N} {\left( -1 \right)^n \over n! } \left( {\lambda \over 4} x^4 \right)^n e^{- x^2 } \\ \label{exint3} & = & \pi^{-1/2} \sum_{n=0}^{N} \int_{- x_{c,N} }^{x_{c,N}} {\rm d} x {\left( -1 \right)^n \over n! } \left( {\lambda \over 4} x^4 \right)^n e^{- x^2 } \end{eqnarray} will converge to the desired integral \begin{equation}\label{desint} z (\lambda) = \pi^{-1/2} \int_{- \infty}^{\infty} {\rm d} x e^{ - \left( x^2 + {\lambda \over 4} x^4 \right) }. \end{equation} In (\ref{exint2}) the change in the limits of integration from $\pm \infty$ to $\pm x_{c,N}$ is just due to the definition of $f'_N$ in Eq.(\ref{convseq}). The interchange between sum and integral in (\ref{exint3}) is now allowed because in the region $\left[ - x_{c,N}, x_{c,N} \right]$ we have uniform convergence (this is a stronger condition than dominated convergence). The resulting integrals are not Gaussian due to the finite limits of integration. We will show how they can be calculated using only Gaussian integrals. A trivial way to achieve convergence of the sequence of integrals of the $f'_N$ of Eq.~(\ref{convseq}) towards~(\ref{desint}) amounts to keep $x_{c,N}$ equal to a finite constant $``a"$ independent of N, while taking the limit $N \rightarrow \infty$. In this limit, Eq.~(\ref{exint3}) becomes identical to $\pi^{-1/2} \int_{-a}^{a} {\rm d} x e^{ - \left( x^2 + {\lambda \over 4} x^4 \right) }$, since for finite $a$ the Taylor series of the integrands converge uniformly. Therefore, as already said, the interchange between sum and integral is legal. Finally, taking the limit $a \rightarrow \infty$, we would obtain the desired convergence towards $z (\lambda)$. However, better use can be made of the information available in $f'_N$ for finite $N$. For example, for every finite $N$, we can choose $x_{c,N}$ so that \begin{equation}\label{diff} |f'_N (x) - f (x) | \le {\epsilon_{T,N} \over 2 x_{c,N}} \quad {\rm for} \quad |x| < x_{c,N}, \end{equation} with $\epsilon_{T,N}$ going to zero as $N \rightarrow \infty$. Then, since we have \begin{equation}\label{diff2} |f'_N (x) - f (x) | \le e^{- \left( x_{c,N}^2 + {\lambda \over 4} x_{c,N}^4 \right)} \equiv {\epsilon_{c,N} \over 2} \quad {\rm for} \quad |x| > x_{c,N}, \end{equation} the $f'_N (x)$ will uniformly converge towards the exact integrand $f(x)$ if~(\ref{diff}) is consistent with $x_{c,N} \rightarrow \infty$ when $N \rightarrow \infty$. Indeed, if this happens, we would have \begin{equation} | \int_{- \infty}^{\infty} \left( f(x) - f_N (x) \right) {\rm d} x | \le \epsilon_{T,N} + \epsilon_{c,N} \rightarrow 0\quad {\rm when} \quad N \rightarrow \infty. \end{equation} The term $\epsilon_{T,N}$ comes trivially from~(\ref{diff}), while $\epsilon_{c,N}$ comes from~Eq.(\ref{diff2}) and the inequality \begin{equation}\label{esterr} \int_{ x_{c,N} }^{\infty} e^{- \left( x^2 + {\lambda \over 4} x^4 \right)} {\rm d} x \le e^{- \left( x_{c,N}^2 + {\lambda \over 4} x_{c,N}^4 \right)} = \epsilon_{c,N}, \end{equation} valid for $x_{c,N}>1$. Applying Taylor's theorem to the function $e^{- \lambda x^4 / 4}$ one can easily show that the condition~(\ref{diff}) is satisfied if \begin{equation}\label{xcNasfunctN} x_{c,N} = \left[ \left( N+1 \right) ! {\epsilon_{T,N} \over 2 } \left( {4 \over \lambda} \right)^{(N+1)} \right]^{1/(4(N + 5/4))}. \end{equation} Note that the non-analytic dependence of $x_{c,N}$ on $\lambda$ arises automatically from the imposition of Eq.~(\ref{diff}) to satisfy the hypothesis of the Lebesgue's theorem. Remember that the only condition on $\epsilon_{T,N}$ to achieve convergence of the sequence of integrals is to go to zero when $N \rightarrow \infty$ consistently with $x_{c,N} \rightarrow \infty$ in that limit. Choosing for example \begin{equation}\label{prescrerr} \epsilon_{T,N} = e^{- 4 N^{1/4}}, \end{equation} we obtain asymptotically, \begin{equation}\label{xcNassymp} x_{c,N} \rightarrow \left( 4 N / e \lambda \right)^{1/4}. \end{equation} This implies (through Eq.~(\ref{diff2})), \begin{equation}\label{calcerr} \epsilon_{c,N} \rightarrow e^{- \left( 4 N / e \lambda \right)^{1/2} - N / e}. \end{equation} Equations~(\ref{prescrerr}) and~(\ref{calcerr}) show the exponential rate at which the convergence of the sequence of integrals take place. Clearly the form~(\ref{prescrerr}) for $\epsilon_{T,N}$ is not unique, not even the most efficient one, but enough to achieve convergence. In the table~(\ref{table}) one can appreciate the numerical convergence \begin{table}[t]\caption{ Integration over the small field configurations only produces a convergent series. In the last column the improvement over the perturbative values can be appreciated.\label{table}} \begin{tabular}{|c|c|c|c|} \hline Order & Exact value ($\lambda=4/10$) & Conv. series & Pert. series \\ \hline 2 & 0.837043 & 0.803160 & 0.848839 \\ \hline 4 & 0.837043 & 0.830264 & 0.854087 \\ \hline 6 & 0.837043 & 0.835516 & 0.901897 \\ \hline 8 & 0.837043 & 0.836667 & 1.316407 \\ \hline 20 & 0.837043 & 0.837044 & 2.33755 $ 10^8$ \\ \hline \end{tabular} \end{table} for $\lambda = 4/10$. \\ Up to now we have proved that the general strategy of section~\ref{sec:converg} does, in fact, generate a convergent sequence towards $z (\lambda)$. However, the resulting integrals in~(\ref{exint3}) are not Gaussians, making the applicability of the method in Quantum Field Theory dubious, to say the least. We will show now that the integrals of Eq~(\ref{exint3}) can be computed, using Eq.~(\ref{win}) with $u = (x/x_{c,N})^2$, calculating only Gaussian integrals. The steps involved are \begin{eqnarray}\label{calcnonGaussint1} \int_{-x_{c,N}}^{x_{c,N}} x^r e^{-x^2} {\rm d} x &=& \int_{- \infty}^{\infty} x^r e^{-x^2} \lim_{M \rightarrow \infty} W(M,x,x_{c,N}) {\rm d} x \\ \label{calcnonGaussint2} &=& \lim_{M \rightarrow \infty} \int_{-\infty}^{\infty} x^r e^{-x^2} W(M,x,x_{c,N}) {\rm d} x \\ &=& \lim_{M \rightarrow \infty} \sum_{n=0}^{M} {1 \over n!} \left( {M \over x_{c,N}^2} \right)^n \int_{-\infty}^{\infty} e^{- \left( 1 + M / x_{c,N}^2 \right) x^2} x^{2n + r} {\rm d} x \nonumber \\ \label{calcnonGaussint3} \end{eqnarray} The two properties of $W$ validate both equality~(\ref{calcnonGaussint1}) and (because of the uniformity of the convergence in $W$)~(\ref{calcnonGaussint2}). In the last line~(\ref{calcnonGaussint3}) we just make explicit the meaning of (\ref{calcnonGaussint2}). So it is clear that these two properties are enough to prove the validity of~(\ref{calcnonGaussint3}), where only Gaussian integrals are present. But it is a good exercise to find a {\it direct} proof of it in the case at hand, where everything can be computed exactly. We do this next. For $r$ odd the integrals vanish, so let's consider the case $r$ even, that is, $r = 2 t$, for any integer $t$. On the one hand we have \begin{equation}\label{left} \int_{-x_{c,N}}^{x_{c,N}} x^{2t} e^{-x^2} {\rm d} x = \left( x_{c,N} \right)^{2t+1} \sum_{k=0}^{\infty} {(-1)^k \over k!} {\left( x_{c,N} \right)^{2k} \over \left( k+t+ 1/2 \right)} \end{equation} where the necessary interchange between sum and integral to arrive to the result is allowed due to the uniform convergence of the Taylor series of $e^{- x^2}$ in the finite segment~$\left[ -x_{c,N}, x_{c,N} \right]$. On the other hand, \begin{eqnarray}\label{right1} & & \lim_{M \rightarrow \infty} \sum_{n=0}^{M} {1 \over n!} \left( {M \over x_{c,N}^2} \right)^n \int_{-\infty}^{\infty} e^{- \left( 1 + M / x_{c,N}^2 \right) x^2} x^{2 \left( n + t \right)} {\rm d} x \qquad\qquad\qquad \\ \label{right2} &=& \lim_{M \rightarrow \infty} \sum_{n=0}^{M} {1 \over n!} \Gamma \left( n+t+ 1/2 \right) \left( {x_{c,N}^2 \over M} \right)^{t + 1/2} \left(1 + {x_{c,N}^2 \over M} \right)^{- \left( n +t + 1/2 \right)} \\ \label{right3} &=& \lim_{M \rightarrow \infty} \sum_{n=0}^{M} {1 \over n!} \left( {x_{c,N}^2 \over M} \right)^{t + 1/2} \sum_{k=0}^{\infty} {(-1)^k \over k!} \Gamma \left( n+t+k+ 1/2 \right) \left( {x_{c,N}^2 \over M} \right)^{k} \\ &=& \left( x_{c,N} \right)^{2t+1} \sum_{k=0}^{\infty} {(-1)^k \over k!} {\left( x_{c,N} \right)^{2k} \over \left( k+t+ 1/2 \right)} \left[ \lim_{M \rightarrow \infty} {(k+t+ 1/2) \over M^{(k+t+ 1/2)}} \sum_{n=0}^{M} { \Gamma \left( n+t+k+ 1/2 \right) \over n!} \right] \nonumber \\ \label{right4} \end{eqnarray} In line~(\ref{right2}) we have used the equation \begin{equation}\label{innt} \int_{- \infty}^{\infty} x^{2n} e^{-p x^2} {\rm d} x = {\Gamma (n + 1/2) \over p^{n+ 1/2}}, \end{equation} in line (\ref{right3}) we have expanded the last term of (\ref{right2}) in powers of $x_{c,N}^2 / M$ and carried out some cancellations, and finally in (\ref{right4}) we have interchanged the $M \rightarrow \infty$ limit with the infinite sum in $k$. Comparing (\ref{left}) and (\ref{right4}), we see that the validity of Eq. (\ref{calcnonGaussint3}) depends on the validity of the equation \begin{equation}\label{identity} \lim_{M \rightarrow \infty} {(k+t+ 1/2) \over M^{(k+t+ 1/2)}} \sum_{n=0}^{M} { \Gamma \left( n+k+t+ 1/2 \right) \over n!} = 1\quad \forall \ {\rm integers} \ k, t > 0 \end{equation} That this identity holds for every integer $t$ and $k$ can be seen by considering the following analytic function of the complex variable $z$: \begin{equation}\label{analfunct} O (z) \equiv \lim_{M \rightarrow \infty} {(1/z) \over M^{(1/z)}} \sum_{n=0}^{M} { \Gamma \left( n+1/z \right) \over \Gamma \left( n+1 \right)!}. \end{equation} If the identities (\ref{identity}) hold, this function must be identically 1, since for $1/z_j = j + 1/2 $ with $j$ integer it reduces to them, and for ever increasing $j$, we obtain a sequence accumulating at $z=0$ on which the function should be 1. Conversely we will prove that $O(z)$ is indeed identically 1 as an analytic function of $z$, proving in consequence the identities (\ref{identity}) for arbitrary $t$ and $k$. Consider the sequence $1/z_j = j + 1$ for $j$ integer. This sequence also accumulates at $z=0$, and for all its points we have \begin{eqnarray}\label{proofident} O \left( 1/ ( j + 1) \right) &=& \lim_{M \rightarrow \infty} {(j+1) \over M^{(j+1)}} \sum_{n=0}^{M} { \Gamma \left( n+ j+1 \right) \over \Gamma \left( n+1 \right)!} \\ &=& \lim_{M \rightarrow \infty} {(j+1) \over M^{(j+1)}} \sum_{n=0}^{M} \Pi_{i=1}^j (i+n) \\ &=& \lim_{M \rightarrow \infty} {(j+1) \over M^{(j+1)}} \left[ \sum_{n=0}^{M} n^j + {\cal O} (n^{j-1}) \right] \\ &=& \lim_{M \rightarrow \infty} {(j+1) \over M^{(j+1)}} \left[ { M^{(j+1)} \over (j+1)} + {\cal O} (M^j) \right] \stackrel{M \rightarrow \infty}{\longrightarrow} 1 \end{eqnarray} Therefore $O(z) = 1$ for all $z$. This finishes the direct proof of Eq.~(\ref{calcnonGaussint3}). As was mentioned before, Eq.~(\ref{xcNasfunctN}), that was derived independently of any analyticity consideration, and only with the purpose of satisfying the hypothesis of Lebesgue's theorem, introduces a non-analyticity in the sequence of integrals of $f_N'$ even for finite $N$. But even for the case where $x_{c,N}$ is fixed to a constant $a$, discussed before, in which the limit $N \rightarrow \infty$ is taken first, and then $a$ is sent to infinity, and therefore the sequence is made out of truly analytic functions, the convergence towards $z (\lambda)$ is perfectly compatible with analyticity considerations. The functions $\pi^{-1/2} \int_{-a}^{a} {\rm d} x e^{ - \left( x^2 + {\lambda \over 4} x^4 \right) }$ (the result of the $N \rightarrow \infty$ limit), are clearly analytic in $\lambda$. But they converge to (in fact they define!) the nonanalytic function $z (\lambda)$ when $a \rightarrow \infty$. The limit of an infinite sequence of analytic functions does not have to be analytic. Another important issue is that the same method works also for the ``negative mass case", where the Borel resummation method fails. Indeed, from the discussion of this section it must be obvious that, with a proper scaling of $x_{c,N}$, the $f_N'$'s with negative quadratic part of the exponent also converges uniformly towards $e^{ (x^2 - {\lambda \over 4} x^4) }$ for $x$ in $\left[ - x_{c,N} , x_{c,N} \right]$. Therefore, the sequence of integrals is also convergent.
proofpile-arXiv_065-496
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{INTRODUCTION} The laser interferometers like LIGO\cite{LIGO}, VIRGO\cite{VIRGO}, GEO\cite{GEO} and TAMA\cite{TAMA} are currently constructed and the detection of gravitational waves is expected at the end of this century. One of the most important sources of gravitational waves for these detectors is coalescing binary compact stars such as NS-NS, NS-BH and BH-BH binaries. Each mass, each spin and the distance of the binary can be determined by applying the matched filter techniques\cite{CF} to the gravitational wave form of the so-called last three minutes of the binary\cite{C}. In the end of the last three minutes two compact stars coalesce and nonlinear character of the gravity and the tidal effects become important, which will be the most exciting part of the coalescing event. It is known from the study of orbits of a test particle in the Schwarzschild metric that the innermost stable circular orbit (ISCO) exists at $r=6M$. For binary cases, Kidder, Will, \& Wiseman\cite{KWW} investigated a point-mass binary using second post-Newtonian equation of motion, and found that the ISCO of the comparable mass binary is at $r_{gr} \simeq 7M_{tot}$ where $r_{gr}$ is the separation of the binary in the Schwarzschild like radial coordinate. As for the tidal effects, Chandrasekhar\cite{Ch69} studied the Roche limit by using Newtonian gravity and treating binary systems as incompressible homogeneous ellipsoids. He found that the stars are tidally disrupted before contact at $r_t=(2.25M_{tot}/\rho)^{1/3}$ for equal mass binary where $\rho$ is the constant density of ellipsoids. For the typical neutron star with $M_{tot}=2.8M_\odot$ and $\rho =1 \times 10^{15}{\rm g/cm^3}$, $r_t$ becomes $5.6M_{tot}$. Recently Lai, Rasio, \& Shapiro\cite{LRS1}\cite{LRS2} have investigated the binary consist of finite size compressible stars using approximate equilibria. In a series of their papers, they took into account the effect of the quadrupole order deviation of stars and found that the hydrodynamic instability occurs at $r_h=6\sim 7.5 M_{tot}$ for $n=0.5$ polytropic neutron star with the radius $R_0=2.5M_{tot}$. In order to know at what radius the final merging phase begins, all three effects, that is, the general relativistic, tidal and hydrodynamic effects should be taken into account simultaneously since $r_{gr}, r_t$ and $r_h$ have similar values of $\sim 7M_{tot}$. For this purpose we must solve the fully general relativistic equations, which is a difficult 3D problem in numerical relativity\cite{N} although the partly general relativistic results have already been presented \cite{WM} \cite{Shibata}. To know the qualitative feature of the ISCOs we will use various approximations to calculate equilibria of the binary in this paper. The reward is that all the calculations can be done analytically so that our results will contribute some physical aspects to the final understanding of the ISCO. In this paper, we will solve the Roche problem\cite{Ch69} as a model of the binary that consists of a finite size star and a point-like gravity source. A gravitational potential which has unstable circular orbits will be used as an interaction potential of the binary. Specifically, we generalize the pseudo-Newtonian potential proposed by Paczy$\acute{{\rm n}}$sky \& Wiita\cite{PW} to mimic the general relativistic effects. We think that this model mimics a neutron star-black hole binary, and describes its qualitative behavior. This paper is organized as follows. In \S2 the basic equations necessary for constructing equilibrium configurations of the Roche ellipsoids (REs) and the Roche-Riemann ellipsoids (RREs)\cite{Aizenman} are derived in the case that the interaction potential is general. In \S3 circular orbits of the neutron star-black hole binary are calculated using the generalized pseudo-Newtonian potential to mimic the effects of general relativity and the results are shown in \S4. In \S5 our results are compared with those using the Newtonian potential and the second order post-Newtonian equation of motion . We use the units of $c=G=1$ through this paper. \section{GENERALIZED ROCHE-RIEMANN ELLIPSOIDS} We first regard the black hole as a point particle of mass $m_2$ and denote the gravitational potential by the black hole as $V_2(r)$. To mimic the effects of general relativity by $V_2(r)$, we do not fix the form of $V_2(r)$ at the moment. Next we treat the neutron star as an incompressible, homogeneous ellipsoid of the semi-axes $a_1, a_2$ and $a_3$, mass $m_1$ and the density $\rho_1$. We treat the self-gravity of the neutron star ($V_1$) as Newtonian. Although for mathematical convenience we assume the incompressibility for the equation of state, the effect of the compressibility can be easily taken into account in an approximate way\cite{LRS1}. Following Chandrasekhar\cite{Ch69}, we call the neutron star the primary and the black hole the secondary. \subsection{Second Virial Equations} We use the tensor virial method\cite{Ch69}, and derive the equations necessary for constructing equilibrium figures of this system. Choose a coordinate system such that the origin is at the center of mass of the primary, the $x_1$-axis points to the center of mass of the secondary, and $x_3$-axis coincides with the direction of the angular velocity of the binary ${\bf \Omega}$. In the frame of reference rotating with ${\bf \Omega}$, the Euler equations of the primary are written as \begin{eqnarray} \rho_1 \frac{du_{i}}{dt} = -\frac{\partial P}{\partial x_{i}} + \rho_1 \frac{\partial}{\partial x_{i}} \left[ V_{1} + V_{2} + \frac{1}{2} {\Omega}^2 \left\{ {\left( \frac{m_{2} R}{m_{1}+m_{2}} - x_1 \right)}^2 + x_2^2 \right\} \right] + 2 \rho_1 \Omega \epsilon_{il3} u_{l}, \label{eom} \end{eqnarray} where $\rho_1, u_i, P$ and $R$ are the density, the internal velocity, the pressure and the separation between the neutron star and the black hole, respectively. Let us expand the interaction potential $V_2$ in power series of $x_k$ up to the second order. This approximation is justified if $R$ is much larger than $a_1, a_2$ and $a_3$. We assume that the potential $V_2$ is spherically symmetric so that it depends only on the distance $r$ from the center of mass of the secondary as \begin{eqnarray} V_2 = V_2(r), \end{eqnarray} where $r$ is given by \begin{eqnarray} r=\left\{(R-x_1)^2 +x_2^2 +x_3^2\right\}^{1/2}. \end{eqnarray} The expansion of $ V_2(r)$ becomes \begin{eqnarray} V_2 = (V_2)_0 - \left( \frac{\partial V_2}{\partial r} \right)_0 x_1 + \frac{1}{2} \left( \frac{\partial^2 V_2}{\partial r^2} \right)_0 x_1^2 + \frac{1}{2R} \left( \frac{\partial V_2}{\partial r} \right)_0 \left( x_2^2 + x_3^2 \right), \label{V2} \end{eqnarray} where the subscript $0$ denotes the derivatives at the origin of the coordinates. In the case of the circular orbit, we have from the force balance at the center \begin{eqnarray} \frac{m_2 R}{m_1+m_2} {\Omega}^2 = -{\left( \frac{\partial V_2} {\partial r} \right)}_0 (1+\delta), \label{omefo} \end{eqnarray} where $\delta$ is the quadrupole term of the interaction potential\cite{LRS1}. Substituting Eqs.(\ref{V2}) and (\ref{omefo}) into Eq.(\ref{eom}), we have \begin{eqnarray} \rho_1 \frac{du_{i}}{dt} = -\frac{\partial P}{\partial x_{i}} &+& \rho_1 \frac{\partial}{\partial x_{i}} \left[ V_{1} + \delta \left( \frac{\partial V_2}{\partial r} \right)_0 x_1+ \frac{1}{2} {\Omega}^2 \left( x_1^2 + x_2^2 \right) + \frac{1}{2} \left( \frac{{\partial}^2 V_2} {\partial r^2} \right)_0 x_1^2 \right. \nonumber \\ &+& \left. \frac{1}{2R} \left( \frac{\partial V_2}{\partial r} \right)_0 \left( x_2^2 + x_3^2 \right) \right] + 2 \rho_1 \Omega \epsilon_{il3} u_l. \label{feom} \end{eqnarray} Multiplying $x_j$ to Eq.(\ref{feom}) and integrating over the volume of the primary, we have \begin{eqnarray} \frac{d}{dt} \int \rho_1 u_i x_j d^3 {\bf x} &=& 2 T_{ij} + W_{ij} + \left\{ {\Omega}^2 + \left( \frac{{\partial}^2 V_2}{\partial r^2} \right)_0 \right\} \delta_{1i} I_{1j} \nonumber \\ & & + \left\{ {\Omega}^2 + \frac{1}{R} \left( \frac{{\partial} V_2}{\partial r} \right)_0 \right\} \delta_{2i} I_{2j} + \frac{1}{R} \left( \frac{{\partial} V_2} {\partial r} \right)_0 \delta_{3i} I_{3j} \nonumber \\ & & + 2 \Omega \epsilon_{il3} \int \rho_1 u_l x_j d^3 {\bf x} + \delta_{ij} \Pi, \label{sve} \end{eqnarray} where \begin{eqnarray} T_{ij} &\equiv& \frac{1}{2} \int \rho_1 u_{i} u_{j} d^3 {\bf x} : {\rm Kinetic~Energy~Tensor}, \\ W_{ij} &\equiv& \int \rho_1 \frac{\partial V_1}{\partial x_i} x_{j} d^3 {\bf x} : {\rm Potential~Energy~Tensor}, \\ I_{ij} &\equiv& \int \rho_1 x_{i} x_{j} d^3 {\bf x} ~~~: {\rm Moment~of~Inertia~Tensor}, \end{eqnarray} and \begin{eqnarray} \Pi &\equiv& \int P d^3 {\bf x}. \end{eqnarray} In Eq.(\ref{sve}) there is no terms related to $\delta$. Since it is possible to take the coordinate system comoving with the center of mass of the binary system, the term proportional to $\delta$ in Eq.(\ref{feom}) vanishes when we integrate over the volume of the primary. Eq.(\ref{sve}) is the basic equation to construct the equilibrium figures of the Roche ellipsoids (REs) and the Roche-Riemann ellipsoids (RREs) for the general potential $V_2(r)$. \subsection{Equilibrium Roche-Riemann Sequence} In this subsection, we show how to construct the equilibrium figures of the RREs. The Roche-Riemann ellipsoid is the equilibrium in which the shape of the primary does not change in the rotating frame although the uniform vorticity exists inside the primary. We restrict the problem to the simplest case where the uniform vorticity of the primary is parallel to the rotation axis, i.e. the primary is the Riemann S-type ellipsoid\cite{Ch69}. We set the coordinate axes to coincide with the principal axes of the primary. For the uniform vorticity $\zeta$, the internal velocity $u_i$ in the rotating frame is given by \begin{eqnarray} u_1 &=& Q_1 x_2, \\ u_2 &=& Q_2 x_1, \end{eqnarray} and \begin{eqnarray} u_3 &=& 0, \end{eqnarray} where \begin{eqnarray} Q_1 &=& - \frac{a_1^2}{a_1^2 + a_2^2} \zeta \end{eqnarray} and \begin{eqnarray} Q_2 &=& \frac{a_2^2}{a_1^2 + a_2^2} \zeta. \end{eqnarray} For the stationary equilibrium, Eq.(\ref{sve}) is rewritten as \begin{eqnarray} Q_{ik}Q_{jl}I_{kl} + W_{ij} &+& \left\{ {\Omega}^2 + \left( \frac{{\partial}^2 V_2}{\partial r^2} \right)_0 \right\} \delta_{1i} I_{1j} + \left\{ {\Omega}^2 + \frac{1}{R} \left( \frac{{\partial} V_2}{\partial r} \right)_0 \right\} \delta_{2i} I_{2j} \nonumber \\ &+& \frac{1}{R} \left( \frac{{\partial} V_2}{\partial r} \right)_0 \delta_{3i} I_{3j} + 2 \Omega \epsilon_{il3} Q_{lk} I_{kj} = - \delta_{ij} \Pi, \label{stasve} \end{eqnarray} where $Q_{ij}$ is not zero only for \begin{eqnarray} Q_{12} &=& Q_1, \\ Q_{21} &=& Q_2. \end{eqnarray} Eq.(\ref{stasve}) has only diagonal components as \begin{eqnarray} Q_1^2 I_{22} + W_{11} + \left\{ {\Omega}^2 + \left( \frac{{\partial}^2 V_2} {\partial r^2} \right)_0 \right\} I_{11} + 2 \Omega Q_2 I_{11} &=& - \Pi, \label{xsve} \\ Q_2^2 I_{11} + W_{22} + \left\{ {\Omega}^2 + \frac{1}{R} \left( \frac{{\partial} V_2} {\partial r} \right)_0 \right\} I_{22} - 2 \Omega Q_1 I_{22} &=& - \Pi, \label{ysve} \end{eqnarray} and \begin{eqnarray} W_{33} + \frac{1}{R} \left( \frac{{\partial} V_2} {\partial r} \right)_0 I_{33} &=& - \Pi. \label{zsve} \end{eqnarray} We assume for simplicity that the gravitational potential of the primary is Newtonian. In this case, the potential energy tensor and the moment of inertia tensor of the incompressible, homogeneous ellipsoids are calculated as \begin{eqnarray} W_{ij} &=& -2 \pi \rho_1 A_{i} I_{ij}, \end{eqnarray} and \begin{eqnarray} I_{ij} &=& \frac{1}{5} m_1 a_{i}^2 \delta_{ij}, \end{eqnarray} where \begin{eqnarray} A_{i} &=& a_{1} a_{2} a_{3} \int_0^{\infty} \frac{du} {\Delta \left(a_{i}^2 +u \right)} , \end{eqnarray} and \begin{eqnarray} \Delta^2 &=& \left( a_{1}^2 +u \right) \left( a_{2}^2 +u \right) \left( a_{3}^2 +u \right). \end{eqnarray} Eliminating $\Pi$ from Eqs.(\ref{xsve})-(\ref{zsve}), we have \begin{eqnarray} \left[ \left\{ 1+ 2\frac{a_2^2}{a_1^2+a_2^2} f_R + \left( \frac{a_1a_2}{a_1^2+a_2^2} f_R \right)^2 \right\} \Omega^2 + \left( \frac{\partial^2 V_2}{\partial r^2} \right)_0 \right] a_{1}^2 &-& \frac{1}{R} \left( \frac{{\partial} V_2}{\partial r} \right)_0 a_{3}^2 \nonumber \\ &=& 2 \pi \rho_1 \left( a_{1}^2 - a_{3}^2 \right) B_{13} \label{sve1} \end{eqnarray} and \begin{eqnarray} \left[ \left\{ 1+ 2\frac{a_1^2}{a_1^2+a_2^2} f_R + \left( \frac{a_1a_2}{a_1^2+a_2^2} f_R \right)^2 \right\} \Omega^2 + \frac{1}{R} \left( \frac{\partial V_2}{\partial r} \right)_0 \right] a_{2}^2 &-& \frac{1}{R} \left( \frac{\partial V_2} {\partial r} \right)_0 a_{3}^2 \nonumber \\ &=& 2 \pi \rho_1 \left( a_{2}^2 - a_{3}^2 \right) B_{23}, \label{sve2} \end{eqnarray} where \begin{eqnarray} f_{R} \equiv \frac{\zeta}{\Omega} \end{eqnarray} and the following relations are used; \begin{eqnarray} a_i^2 A_i -a_j^2 A_j &=& \left( a_i^2-a_j^2 \right) B_{ij}, \end{eqnarray} where \begin{eqnarray} B_{ij} &=&a_1 a_2 a_3 \int_0^{\infty} \frac{u du} {\Delta (a_i^2 +u)(a_j^2 +u)}. \end{eqnarray} Now from Eq.(2.5) $\Omega$ is given by \begin{eqnarray} \Omega^2 = -\frac{1+p}{R} \left( \frac{\partial V_2}{\partial r} \right)_0 (1+\delta) ~~~~~~~~~~\left( p \equiv \frac{m_1}{m_2} \right). \label{omega} \end{eqnarray} Dividing Eq.(\ref{sve1}) by Eq.(\ref{sve2}), we have the equation to determine the Roche-Riemann sequences as \begin{eqnarray} \frac{\left[ (1+p)(1+\delta) \left\{ 1+ 2\frac{a_2^2}{a_1^2+a_2^2} f_R +\left( \frac{a_1 a_2}{a_1^2+a_2^2} f_R \right)^2 \right\} -R \left( \frac{\partial^2 V_2}{\partial r^2} \right)_0/ \left( \frac{\partial V_2}{\partial r} \right)_0 \right]a_1^2 +a_3^2} {\left[ (1+p)(1+\delta) \left\{ 1+2 \frac{a_1^2}{a_1^2+a_2^2} f_R +\left( \frac{a_1a_2}{a_1^2+a_2^2} f_R \right)^2 \right\} -1 \right] a_2^2 + a_3^2} = \frac{\left( a_1^2 - a_3^2 \right) B_{13}} {\left( a_2^2 - a_3^2 \right) B_{23}}. \label{gaxis} \end{eqnarray} Using Eqs.(\ref{sve2}) and (\ref{omega}), we can determine the orbital angular velocity $\Omega$ by \begin{eqnarray} \frac{\Omega^2}{\pi \rho_1}=\frac{2(1+p)(1+\delta)\left(a_2^2 -a_3^2 \right) B_{23}}{ \left[ (1+p)(1+\delta)\left\{ 1+ 2 \frac{a_1^2}{a_1^2+a_2^2} f_R + \left( \frac{a_1a_2}{a_1^2+a_2^2} f_R \right)^2 \right\}-1 \right] a_2^2 + a_3^2}. \label{gOme} \end{eqnarray} Note that $f_R$ is related to the circulation ${\cal C}$ as \begin{eqnarray} {\cal C} &=& \oint {\bf u}_{inertial} \cdot d{\bf l}= \pi a_1 a_2 (2 + f_R) \Omega, \end{eqnarray} where \begin{eqnarray} {\bf u}_{inertial} &=& \left\{ \begin{array}{@{\,}ll} (u_{inertial})_1 = (Q_1 -\Omega) x_2, \\ (u_{inertial})_2 = (Q_2 + \Omega) x_1-\frac{R}{1+p} \Omega~ \\ (u_{inertial})_3 = 0. \end{array} \right. \end{eqnarray} If there is no viscosity inside the primary, the circulation should be conserved from Kelvin's circulation theorem. \subsection{Total Angular Momentum} The total energy and the total angular momentum of the binary are the decreasing functions of time since the gravitational waves are emitted. If the total angular momentum has its minimum at some separation of the binary, we regard this point as the ISCO \footnote{Lai, Rasio, \& Shapiro show in appendix D of \cite{LRS1} that the true minimum point of the total energy coincides with that of the total angular momentum. Strictly speaking, if the rotation includes only to the quadrupole order, this coincidence fails. However the difference is as small as the numerical accuracy \cite{LRS1}.}. The total angular momentum of our system which is the sum of the orbital and the spin angular momentum is given by \begin{eqnarray} J_{tot} &=& m_1 r_{cm}^2 \Omega + m_2 (R-r_{cm})^2 \Omega + I \Omega + \frac{2}{5}m_1 \frac{a_1^2 a_2^2} {a_1^2 + a_2^2} \zeta \nonumber \\ &=& \frac{m_1 m_2}{m_1 + m_2} R^2 \Omega \left\{ 1 + \frac{1}{5} (1+p) \frac{1}{R^2} \left( a_1^2 + a_2^2 + 2\frac{a_1^2 a_2^2}{a_1^2+a_2^2} f_R \right) \right\} \label{angmom} \end{eqnarray} where \begin{eqnarray} r_{cm} = \frac{m_{2} R}{m_1+m_2}. \end{eqnarray} The first term in the braces of the right hand side of Eq.(\ref{angmom}) comes from the orbital angular momentum of the binary system and the second does from the spin angular momentum of the primary. \section{GENERALIZED PSEUDO-NEWTONIAN POTENTIAL} There are variety of choices of $V_2(r)$ to mimic the general relativistic effects of the gravitation. We generalize the so-called pseudo-Newtonian potential proposed by Paczy$\acute{{\rm n}}$sky \& Wiita\cite{PW} originally. This potential fits the effective potential of the Schwarzschild black hole quite well as we will show later. We will use the generalized pseudo-Newtonian potential defined by \begin{eqnarray} V_2(r) &=& \frac{m_2}{r-r_{pseudo}} \label{pnpot},\\ r_{pseudo} &=& r_s \left\{ 1+ g(p) \right\}, \\ g(p) &=& \frac{7.49p}{6(1+p)^2} - \frac{10.4 p^2}{3(1+p)^4} + \frac{29.3 p^3}{6(1+p)^6}, \label{gp} \\ r_s &\equiv& \frac{2GM_{tot}}{c^2}, \\ M_{tot} &=& m_1 + m_2, \end{eqnarray} where $p=m_1/m_2$ and $g(p)$ is the special term to fit the ISCOs of the hybrid second post-Newtonian calculations by Kidder, Will, \& Wiseman\cite{KWW}. For $p=0$, the generalized pseudo-Newtonian potential agrees with the pseudo-Newtonian potential proposed by Paczy$\acute{{\rm n}}$sky \& Wiita\cite{PW}. Fig.1(a) shows effective potentials (solid lines) and locations of circular orbits (dots) in our generalized pseudo-Newtonian potential ( $p=0$ \& $r_{pseudo}=r_s$) and in the Schwarzschild metric. Although by this choice of the parameter ($r_{pseudo}=r_s$), the locations of the ISCOs in the generalized pseudo-Newtonian potential agree with those in the Schwarzschild metric, the angular momenta at the ISCO are different, that is, the angular momentum in the generalized pseudo-Newtonian potential ($J_{pseudo}$) for $p=0$ is $(9/8)^{1/2}$ times larger than that in the Schwarzschild metric ($J_{Sch}$) at the ISCO. Therefore in Fig.1(a) and (b) we compare circular orbits with different angular momentum related as \begin{eqnarray} J_{pseudo} = \left( \frac{9}{8} \right)^{1/2} J_{Sch}. \end{eqnarray} From Fig.1(b) we see that the radii of the circular orbits of the generalized pseudo-Newtonian potential agrees with those of the effective potential around Schwarzschild black hole within 10\% accuracy near the ISCO. This is the reason why we believe that our generalized pseudo-Newtonian potential expresses the effect of general relativity within 10\% or so. Using Eq.(\ref{pnpot}) and Eq.(\ref{omega}), we can rewrite Eq.(\ref{gOme}) as \begin{eqnarray} \frac{p^2 r_s^3 (\bar{a}/m_1)^3}{12(1+p)^3 R(R-r_{pseudo})^2} - \frac{(a_2^2-a_3^2) B_{23}} { \left[ (1+p)(1+\delta) \left\{ 1+ 2 \frac{a_1^2}{a_1^2 + a_2^2} f_R + \left( \frac{a_1a_2}{a_1^2 + a_2^2} f_R \right)^2 \right\} -1 \right] a_2^2 + a_3^2 } = 0, \label{sequence} \end{eqnarray} where $\bar{a}$ is the mean radius of the primary. In the generalized pseudo-Newtonian case, the quadrupole term $\delta$ is written as \begin{eqnarray} \delta=\frac{3}{10} \left\{2a_1^2- \frac{(3R-r_{pseudo})(R-r_{pseudo})}{3R^2} \left(a_2^2+a_3^2 \right) \right\} \frac{1}{(R-r_{pseudo})^2}. \label{quadru} \end{eqnarray} We also have the separation of the binary as \begin{eqnarray} R=\frac{E}{E-2} r_{pseudo}, \label{sob} \end{eqnarray} where \begin{eqnarray} E \equiv &-&(1+p)(1+\delta) \left\{ 1+ 2\frac{a_2^2}{a_1^2+a_2^2} f_R + \left( \frac{a_1 a_2}{a_1^2 + a_2^2} f_R \right)^2 \right\} - \frac{a_3^2}{a_1^2} \nonumber \\ &+& \left\{ \left[ (1+p)(1+\delta) \left\{ 1 + 2\frac{a_1^2}{a_1^2+a_2^2} f_R + \left( \frac{a_1a_2}{a_1^2+a_2^2} f_R \right)^2 \right\} -1 \right] \frac{a_2^2}{a_1^2} + \frac{a_3^2}{a_1^2} \right\} \frac{(a_1^2-a_3^2) B_{13}}{(a_2^2-a_3^2) B_{23}}. \end{eqnarray} For the given mass ratio $p$, mean radius $/bar{a}/m_1$, circulation parameter $f_{R}$ and axial ratio $a_3/a_1$, we can determine the axial ratio $a_2/a_1$ by solving Eq.(\ref{sequence}) with Eqs.(\ref{quadru}) and (\ref{sob}). Using the axial ratios ($a_2/a_1$, $a_3/a_1$), we are able to calculate the orbital angular velocity by Eq.(\ref{gOme}) and the separation of the binary by Eq.(\ref{sob}). The total angular momentum is calculated by Eq.(\ref{angmom}). Finding the minimum of the total angular momentum, we can determine the location of the ISCO. When the viscosity inside the primary is so effective that no internal motion exists, $f_R=0$ in the above equation, i.e. the Roche ellipsoids (REs). While if the primary is in-viscid and ${\cal C}=0$, we have $f_R=-2$, i.e. the irrotational Roche-Riemann ellipsoids (IRREs). Note that if we substitute the Newtonian potential as an interaction potential, Eqs.(\ref{gaxis}) and (\ref{gOme}) agree with the equations derived by Chandrasekhar\cite{Ch69} in the REs ($f_R=0$) case and those by Aizenman\cite{Aizenman} in the RREs case. \section{RESULTS} Kochanek\cite{Kocha} and Bildsten \& Cutler\cite{BC} showed that the internal structure of a coalescing binary neutron star is the irrotational Roche-Riemann ellipsoids (IRREs). However we calculate both the REs and the IRREs for comparison. We show the results in four cases; 1) the REs with $p=1$ or 0.1, 2) the IRREs with $p=1$ or 0.1. Results are given in Table.I to Table.VI, where $\tilde{\Omega}=\Omega/\sqrt{\pi \rho_1}$ represents the normalized orbital angular velocity, and $\tilde{J}$ denotes the normalized total angular momentum defined by \begin{eqnarray} \tilde{J}=\frac{J_{tot}}{m_1 m_2 (r_s/M_{tot})^{1/2}}. \end{eqnarray} $\bar{a}$ is the mean radius of the primary defined by \begin{eqnarray} \bar{a} = \left( \frac{m_1}{\frac{4}{3} \pi \rho_1} \right)^{1/3}. \end{eqnarray} In Tables I to VI, $\dagger$ means the point of the ISCO defined in this paper, and $\ddagger$ does the point of the Roche limit where the Roche limit is defined by the distance of closest approach for equilibrium to be possible\cite{Ch69}. The values in the parentheses show the power of $10$. \subsection{$p=1$ Case} Fig.2(a) shows $\tilde{J}$ as a function of the normalized separation $R/r_s$. In Fig.2(a), thin solid, dotted and dashed lines are the REs with $\bar{a}/m_1$ being $3$, $5$ and $8$, respectively, while thick solid, dotted and dashed lines are the IRREs with $\bar{a}/m_1$ being $3$, $5$ and $8$, respectively. We defined in \S\S II.C that the location of the ISCO is the minimum point of $\tilde{J}$. From Fig.2(a), we see that the separations of the binary at the ISCO in the IRREs case are almost the same as those in the REs case. Fig.2(b) shows the axial ratios $a_2/a_1$ and $a_3/a_1$ as a function of $R/r_s$. The conventions of lines are the same as those in Fig.2(a). The relation among the length of the axes is $a_1>a_2>a_3$ in the REs, while $a_1>a_3>a_2$ in the IRREs. In the REs, the tidal force makes the $a_1$ axis long while it does $a_2$ and $a_3$ axes short. The centrifugal force makes $a_1$ and $a_2$ axes long. As a result we have $a_1>a_2>a_3$. In the IRREs, in addition to the above effects, the Coriolis force caused by the internal motion of the primary exists. This makes $a_1$ and $a_2$ axes short, which yields $a_1>a_3>a_2$ . From Fig.2(b), it is found that in the IRREs the star with $\bar{a}/m_1=3$ reaches the ISCO at the point of $a_2/a_1 \simeq 0.912$ and $a_3/a_1 \simeq 0.918$. On the other hand, the star with $\bar{a}/m_1=8$ terminates when $a_2/a_1 \simeq 0.760$ and $a_3/a_1 \simeq 0.792$. This means that the primary with the smaller mean radius reaches the ISCO before the shape of the primary deviates from the sphere considerably. This tendency is the same for the REs. \subsection{ $p=0.1$ Case} Fig.3(a) shows $\tilde{J}$ as a function of $R/r_s$. The conventions are the same as those in Fig.2(a) except $\bar{a}/m_1$ being $3$, $5$ and $8$. From Fig.3(a), we see that all lines with the mean radii in the range of $3 \le \bar{a}/m_1 \le 8$ take their minimum at the points near $R/r_s \sim 3.25$. This value is almost the same as the ISCO by Kidder, Will, \& Wiseman\cite{KWW} for $p=0.1$. This means that when $p$ is much less than $1$ and the mean radius of the primary is less than $8m_1$ which corresponds to 17km for $m_1=1.4M_\odot$, the size of the primary has little effect on the ISCO. Fig.3(b) shows the axial ratios $a_2/a_1$ and $a_3/a_1$ as a function of $R/r_s$. The conventions are the same as those in Fig.3(a). We see that the stars with smaller mean radii reach the ISCO even when the deviation from the spherical symmetry is very small. Since the binary system enters an unstable circular orbit before the primary is tidally deformed, the tidal effects are not important. We see that even $\bar{a}/m_1\sim 8$ the minimum values $a_2/a_1$ and $a_3/a_1$ are not small ($\sim 0.85$). \section{DISCUSSIONS} In this section, we discuss the differences between the case of the generalized pseudo-Newtonian potential and that of the Newtonian potential, and between the REs and the IRREs. We will also compare our results with other papers. \subsection{The Case of $p=1$} In Fig.4, the separation $R_{ISCO}/r_s$ is shown as a function of $\bar{a}/m_1$. Thick lines and thin lines represent the case of the generalized pseudo-Newtonian potential and that of the Newtonian potential, respectively. Solid lines and dotted lines express the case of the IRREs and the REs, respectively. We see that for the Newtonian potential, $R_{ISCO}/r_s$ increases in proportion to $\bar{a}/m_1$ regardless of types of ellipsoids, while for the generalized pseudo-Newtonian potential, the behavior of $R_{ISCO}/r_s$ changes around $\bar{a}/m_1 \simeq 3.5$. For $\bar{a}/m_1 \mathrel{\mathpalette\Oversim>} 3.5$, $R_{ISCO}/r_s$ increases in proportion to $\bar{a}/m_1$ as in the case of the Newtonian potential. In this region, the tidal effect dominates the system and the effects of the general relativity become less important. In conclusion, for $\bar{a}/m_1 \mathrel{\mathpalette\Oversim>} 3.5$, the tidal effect dominates the stability of the binary system and for $\bar{a}/m_1 \mathrel{\mathpalette\Oversim<} 3.5$, the binary system is dominated by the relativistic effect, i.e. the fact that the interaction potential has an unstable orbit. From Fig.4 it is also found that the location of the ISCOs in the IRREs case is not so different from that in the REs case for the same $\bar{a}/m_1$. The orbital frequency of the IRREs at the ISCO for $m_1=1.4M_{\odot}$ and $\bar{a}/m_1 = 5$ is estimated from $\tilde{\Omega}$ in Table.IV as \begin{eqnarray} \Omega_{\rm ISCO} &=& 495~[{\rm Hz}]. \end{eqnarray} This value is smaller than that in the Newtonian potential case ($\Omega_{\rm ISCO}^{Newton} = 599~[{\rm Hz}]$). \subsection{The Case of $p=0.1$} If we consider the primary as the neutron star of mass $1.4M_{\odot}$, then, the secondary is regarded as the black hole of mass $14M_{\odot}$. $R_{ISCO}/r_s$ is shown as a function of $\bar{a}/m_1$ in Fig.5. For the Newtonian potential, $R_{ISCO}/r_s$ increases in proportion to $\bar{a}/m_1$ like the case of $p=1$. For the generalized pseudo-Newtonian potential, in the range of $\bar{a}/m_1$ of Fig.5, $R_{ISCO}/r_s$ converges to the value $3.25$ obtained by Kidder, Will, \& Wiseman\cite{KWW}. This is because the existence of the unstable orbit in the generalized pseudo-Newtonian potential influences $R_{ISCO}/r_s$, and this effect dominates when the radius of the primary is small. This is clearer for the small mass ratio $p$. Therefore if $p$ is small, for the range of the radius ($3 \le \bar{a}/m_1 \le 8$) relevant to the neutron star, the effect of the neutron star's size is very small. This comes essentially from the fact that the Newtonian estimate of the Roche radius is smaller than the radius of the ISCO. One can estimate the orbital frequency of the IRREs at the ISCO for $m_1=1.4M_{\odot}$ and $\bar{a}/m_1 = 5$ from $\tilde{\Omega}$ in Table.V as \begin{eqnarray} \Omega_{\rm ISCO} &=& 187~[{\rm Hz}]. \end{eqnarray} \subsection{Comparison with Other Works} Kidder, Will, \& Wiseman\cite{KWW} studied the motion of the point-particle binary systems using {\it hybrid Schwarzschild second post-Newtonian equation of motion}, and obtained the separation at the ISCO ($r_{ISCO}$) expressed by \begin{eqnarray} \frac{r_{ISCO}}{M_{tot}} \simeq 6 + 7.49 \eta - 20.8 \eta^2 +29.3 \eta^3 \label{KWWISCO} \end{eqnarray} where \begin{eqnarray} \eta &=& \frac{m_1 m_2}{M_{tot}^2}. \end{eqnarray} Table.VII shows the comparison of Eq.(\ref{KWWISCO}) with our results for $\bar{a}/m_1 = 5$. It is found that for $p=0.1$, the finite size effect is not important because the primary is much lighter than the secondary, so that the results are almost the same as \cite{KWW}. On the other hand, when $p=1$ the finite size of the primary increases the separation at the ISCO by the general relativistic and tidal effects. Lai, Rasio, \& Shapiro\cite{LRS3}, discussed the relativistic effects on the binary system for compressible Darwin ellipsoids using a simple approximate model\cite{LRS4}. The Darwin ellipsoids are the equilibrium configurations constructed by two identical synchronized finite-size stars including the mutual tidal interactions. In their approach the effects of general relativity and the Newtonian tidal interactions for finite-size compressible stars are combined by hand. While we formulated the problem using arbitrary interaction potentials of the secondary for the incompressible primary. We adopted the semi-relativistic potential called the generalized pseudo-Newtonian potential to mimic the general relativistic effects of gravitation. We solved the equilibria of the REs and the IRREs in this potential. Their results and ours are compared in Table.VIII, where $r_m$ expresses the minimum separation obtained by Lai, Rasio, \& Shapiro\cite{LRS3}. From this table, we see that both results agree rather well inspite of different approaches and approximations. \acknowledgments KT would like to thank H. Sato, K. Nakao and M. Shibata for useful discussions and continuous encouragement. This work was in part supported by a Grant-in-Aid for Basic Research of Ministry of Education, Culture, Science and Sports (08NP0801).
proofpile-arXiv_065-497
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{Introduction} {\em Ab-initio} Hartree-Fock (HF) and configuration interaction (CI) methods are standard tools in computational chemistry nowadays and various program packages are available for accurate calculations of properties of atoms and molecules. For solids, HF calculations have become possible, on a broad scale, with the advent of the program package CRYSTAL \cite{CRYSTAL}. However, the problem of an accurate treatment of electron correlation is not fully settled (for a survey see \cite{FuldeBuch}). Although the absolute value of the HF energy is usually much larger than the correlation energy, the correlation energy is very important for energy differences. For example, the O$^-$ ion is not stable at the HF level, and correlations are necessary in order to obtain even qualitative agreement with the experimental result for the electron affinity of oxygen. In solid state physics, NiO is a well known example of a system which is insulating due to correlations. The most widely used method to include correlations in solids is density functional theory (DFT) \cite{DFT}. DFT has also recently become quite popular for a computationally efficient treatment of exchange and correlation in molecules. However, a systematic improvement towards the exact results is currently not possible with DFT. Wave-function-based methods are more suitable for this purpose. In the last years, Quantum Monte-Carlo calculations have been performed for several systems \cite{QMC}. Correlations are included here by multiplying the HF wavefunction with a Jastrow factor. An approach more closely related to quantum chemistry is the Local Ansatz \cite{Stollhoff,FuldeBuch} where judiciously chosen local excitation operators are applied to HF wavefunctions from CRYSTAL calculations. Some years ago, an incremental scheme has been proposed and applied in calculations for semiconductors \cite{StollDiamant,Beate}, for graphite \cite{StollGraphit} and for the valence band of diamond \cite{Juergen}; here information on the effect of local excitations on solid-state properties is drawn from calculations using standard quantum chemical program packages. In a recent paper \cite{DDFS} we showed that this method can be successfully extended to ionic solids; we reported results for the correlation contribution to the cohesive energy of MgO. In the present article, we apply the scheme to the cohesive energy of CaO, as a second example. In addition, we show how correlations affect the lattice constants of MgO and CaO. For these systems, several calculations have been performed at the HF level with the CRYSTAL code \cite{CausaMgO1,DRFASH,MackrodtCaO,CausaZupan,McCarthy,Catti} as well as with inclusion of correlations using DFT \cite{DRFASH,CausaZupan,McCarthy,Pyper}. \section{The Method} The method of increments can be used to build up correlation effects in solids from local correlation contributions which in term may be obtained by transferring results from suitably embedded finite clusters to the infinite crystal. It has been fully described in \cite{StollDiamant,Beate,StollGraphit,DDFS}, and a formal derivation has been given within the framework of the projection technique \cite{TomSchork}. Thus, we will only briefly repeat the main ideas.\\ (a) Starting from self-consistent field (SCF) calculations localized orbitals are generated which are assumed to be similar in the clusters and in the solid. \\ (b) One-body correlation-energy increments are calculated: in our specific case these are the correlation energies $\epsilon (A)$, $\epsilon (B)$, $\epsilon (C)$, ... of localized orbital groups which can be attributed to X$^{2+}$ (X = Mg, Ca) or O$^{2-}$ ions at ionic positions $A$, $B$, $C$, ... Each localized orbital group is correlated separately. \\ (c) Two-body increments are defined as non-additivity corrections: \begin{eqnarray*} \Delta \epsilon {(AB)}=\epsilon (AB) - \epsilon (A) -\epsilon (B), \end{eqnarray*} where $\epsilon {(AB)}$ is the correlation energy of the joint orbital system of $AB$.\\ (d) Three-body increments are defined as \begin{eqnarray*} \Delta\epsilon(ABC) = & \epsilon(ABC) - [\epsilon(A) + \epsilon(B) + \epsilon(C)] - \nonumber \\ & [\Delta\epsilon(AB) + \Delta\epsilon(AC) + \Delta\epsilon(BC)]. \end{eqnarray*} Similar definitions apply to higher-body increments. \\ (e) The correlation energy of the solid can now be expressed as the sum of all possible increments: \begin{equation} \epsilon_{\rm bulk} = \sum_A \epsilon(A) + \frac{1}{2} \sum_{A,B} \Delta\epsilon(AB) + \frac{1}{3!} \sum_{A,B,C} \Delta\epsilon(ABC) + ... \end{equation} Of course, this only makes sense if the incremental expansion is well convergent, i.e. if $\Delta \epsilon (AB)$ rapidly decreases with increasing distance of the ions at position $A$ and $B$ and if the three-body terms are significantly smaller than the two-body ones. A pre-requisite is that the correlation method used for evaluating the increments must be size-extensive: otherwise the two-body increment $\Delta \epsilon (AB)$ for two ions $A$ and $B$ at infinite distance would not vanish. In our present work, we used three different size-extensive approaches, cf. Sect. 2.1. Finally, the increments must be transferable, i.e.\ they should only weakly depend on the cluster chosen. \subsection{Correlation Methods} In this section we want to give a brief description of the correlation methods used. In the averaged coupled-pair functional (ACPF \cite{ACPF}) scheme, the correlation energy is expressed in the form \begin{equation} E_{\rm corr}[\Psi_c]=\frac{< \Psi_{SCF}+\Psi_c|H-E_{SCF} |\Psi_{SCF}+\Psi_c >} {1+g_c<\Psi_c|\Psi_c>} \end{equation} with $\Psi_{SCF}$ being the SCF-wavefunction (usually of the spin-restricted Hartree-Fock type) and $\Psi_c$ the correlation part of the wavefunction, \begin{eqnarray} |\Psi_c>= \sum_{a \atop r}c_{a}^{r}a^+_r a_a|\Psi_{SCF}> + \sum_{a<b \atop r<s}c_{ab}^{rs}a^+_r a^+_s a_a a_b|\Psi_{SCF}>; \end{eqnarray} $g_c$ is chosen as $\frac{2}{n}$ in order to make the expression (2) approximately size-consistent ($n$ being the number of correlated electrons). For more details (and the extension to multi-reference cases), see \cite{ACPF}.\\ In the coupled-cluster singles and doubles (CCSD \cite{CCSD}) scheme, the wavefunction is expressed with the help of an exponential ansatz: \begin{eqnarray} |\Psi_{CCSD}>=\mbox{exp}({\sum_{a \atop r}c_a^r a^+_r a_a + \sum_{a<b \atop r<s}c_{ab}^{rs}a^+_r a^+_s a_a a_b})|\Psi_{SCF}>. \end{eqnarray} $a^+$ ($a$) are creation (annihilation) operators of electrons in orbitals which are occupied ($a$, $b$) or unoccupied ($r,$ $s$) in the SCF wavefunction.\\ Finally, in the CCSD(T) scheme, three particle excitations are included by means of perturbation theory as proposed in \cite{Raghavachari}.\\ We used these three methods to compare their quality in applications to solids. It turns out that ACPF and CCSD give very similar results, while CCSD(T) yields slighty improved energies \cite{CCSDTWerner}. Altogether, the results are not strongly dependent on the methods and no problem arises, therefore, if only one method should be applicable (as is the case for low-spin open-shell systems, where CCSD and CCSD(T) are not yet readily available). All calculations of this work were done by using the program package MOLPRO \cite{KnowlesWerner,MOLPROpapers}. \section{Cohesive Energy of CaO} \subsection{Basis sets and test calculations} For calculating the correlation contribution to the cohesive energy of CaO, we closely follow the approach of Ref.\ \cite{DDFS}. For oxygen we choose a $[5s4p3d2f]$ basis set \cite{Dunning}. Calcium is described by a small-core pseudopotential replacing the $1s$, $2s$ and $2p$ electrons \cite{KauppCa}, and a corresponding $[6s6p5d2f1g]$ valence basis set (from ref. \cite{KauppCa}, augmented with polarization functions $f_1$=0.863492, $f_2$=2.142 and $g$=1.66) is used. All orbitals are correlated, with the exception of the O $1s$ core. In particular, the correlation contribution of the outer-core Ca $3s$ and $3p$ orbitals is explicitly taken into account. We did not use a large-core (X$^{2+}$) pseudopotential and a core polarization potential (CPP) for treating core-valence correlation as was done in the case of MgO \cite{DDFS}, because Ca is close to the transition metals and excitations into $d$-orbitals are important. The influence of the latter on the X$^{2+}$ core cannot be well represented by a CPP since the $3d$ orbitals are core-like themselves (cf.\ the discussion in \cite{MuellerFleschMeyer}). Correlating the Ca outer-core orbitals explicitly, using the small-core (Ca$^{10+}$) pseudopotential, we circumvent this problem. Using this approach, we performed test calculations for the first and second ionization potential of the Ca atom (Table 1) and calculated spectroscopic properties of the CaO molecule (Table 2). In both cases, we obtain good agreement with experiment. \subsection{Intra-atomic Correlation} We first calculated one-body correlation-energy increments. For Ca$^{2+}$, the results are virtually independent of the solid-state surroundings. This was tested by doing calculations for a free $\rm {Ca}^{2+}$ and a $\rm {Ca}^{2+}$ embedded in point charges. (A cube of 7x7x7 ions was simulated by point charges $\pm$2, with charges at the surface planes/edges/corners reduced by factors 2/4/8, respectively.) In the case of $\rm {O}^{2-}$, of course, the solid-state influence is decisive for stability, and we took it into account by using an embedding similar to that of Ref.\ \cite{DDFS}: the Pauli repulsion of the 6 nearest $\rm {Ca}^{2+}$ neighbours was simulated by large-core pseudopotentials \cite{Fuentealba}, while the rest of a cube of 7x7x7 lattice sites was treated in point-charge approximation again. A NaCl-like structure with a lattice constant of 4.81 \AA \mbox{ } was adopted. (The experimental value for the lattice constant is 4.8032 \AA \mbox{ } at a temperature of T=17.9 K \cite{LandoltGitter}). We performed similar calculations for various other finite-cluster approximations of the CaO crystal, in order to insure that the results are not sensitive to lattice extensions beyond the cube mentioned above. The results for the one-body correlation-energy increments are shown in Table 3. It is interesting to note that the absolute value of the Ca $\rightarrow \rm{Ca}^{2+}$ increment is larger than in the case of Mg, although the electron density in the valence region of Ca is lower than for Mg. The larger correlation contribution for Ca can be rationalized by the fact that excitations into low-lying unoccupied $d$-orbitals are much more important for Ca than for Mg. This is a result which would be difficult to explain by density functional theory: in a local-density framework, higher density leads to a higher absolute value of the correlation energy. In Ref.\ \cite{DDFS} we argued that the increment in correlation energy $\epsilon(\rm {embedded}$ $\rm O^{2-})-\epsilon(\rm {free}$ ${\rm O})$ is not just twice the increment $\epsilon(\rm {free}$ $\rm O^{-})-\epsilon(\rm {free}$ $\rm O)$. However, comparing the increments $\epsilon$ (embedded $\rm {O}^{2-})-\epsilon (\rm {embedded} $ $\rm O)$ and $\epsilon$ ( {embedded} $\rm {O}^{-})-\epsilon (\rm embedded $ ${\rm O})$ one finds a factor very close to two. This can be seen from Table 4 where we compare the increments in the case of MgO. Thus, for the embedded species linear scaling is appropriate as in the case of the gas-phase isoelectronic series Ne, Ne$^+$, Ne$^{2+}$: there, the increments in correlation energy are 0.0608 H (Ne$^{2+}$ $\rightarrow$ Ne$^+$) and 0.0652 H (Ne$^{+}$ $\rightarrow$ Ne) \cite{Davidson}. Table 4 also shows that the correlation contribution to the electron affinity of the oxygen atom is {\em smaller} for the embedded species than in the gas phase. This is due to the fact that energy differences to excited-state configurations become larger when enclosing O$^{n-}$ in a solid-state cage. Once again, this is at variance with a LDA description as the electron density in the case of the embedded $\rm O^-$ is more compressed than in the case of a free $\rm O^-$. In Figure 1 we show the charge density distribution of O$^{2-}$, again in the case of MgO. We used basis functions on both O and Mg; the Mg $1s$, $2s$ and $2p$-electrons are replaced by a pseudopotential. One recognizes the minimum near the Mg$^{2+}$ cores, where the Pauli repulsion prevents the oxygen electrons from penetrating into the Mg$^{2+}$ core region. This way, the solid is stabilized. The 6$^{th}$ contour line, counting from Mg to O, is the line which represents a density of 0.002 a.u. This is the density which encloses about 95 \% of the charge and was proposed as an estimate of the size of atoms and molecules \cite{Bader}. The sum of the intra-ionic correlation-energy increments discussed in this subsection turns out to yield only $\sim$ 60 \% of the correlation contribution of the cohesive energy of CaO. This percentage is quite similar to that obtained for MgO \cite{DDFS}, at the same level. Thus, although MgO and CaO are to a very good approximation purely ionic solids, the inter-atomic correlation effects to be dealt with in the next subsection play an important role. \subsection{Two- and three-body increments} When calculating two-body correlation-energy increments, point charges or pseudopotentials surrounding a given ion have to be replaced by 'real' ions. In the case of an additional 'real' $\rm {O}^{2-}$, its next-neighbour shell also has to be replaced by a cage of pseudopotentials simulating $\rm Ca^{2+}$. This way the increments shown in Table 3 are obtained. It turns out that the Ca-O increments are much more important than the O-O increments, while Ca-Ca increments are negligibly small. The changes with respect to MgO \cite{DDFS} can easily be rationalized: On the one hand, the lattice constant is larger than in the case of MgO (4.81 \AA \mbox{ } vs. 4.21 \AA), which reduces the van der Waals interaction and makes the O-O increments smaller. On the other hand, the polarizability of $\rm Ca^{2+}$ is by a factor of more than 6 higher than that of $\rm Mg^{2+}$ (see for example Ref. \cite{Fuentealba}), which leads to large Ca-O increments. We show the van der Waals-like decay in Figure 2 by plotting the two-body increments O-O for CaO from CCSD calculations (without including weight factors). By multiplying with the sixth power of the distance, one can verify the van der Waals-law. Plots for the other two-body increments are qualitatively similar. Three-body increments contribute with less than 2 \% to the correlation piece of the bulk cohesive energy and may safely be neglected, therefore. A survey of the convergency pattern of the incremental expansion, for both CaO and MgO, is given in Figs.\ 3 and 4. \subsection{Sum of increments} Adding up the increments of sections 3.2 and 3.3 (cf.\ Table 5), we obtain between 71 and 78 \% of the 'experimental' correlation contribution to the cohesive energy which we define as the difference of the experimental cohesive energy (11.0 eV, \cite{CRC}) plus the zero-point energy (which is taken into account within the Debye approximation and is of the order 0.1 eV) minus the HF binding energy (7.6 eV, Ref. \cite{MackrodtCaO}). The percentage obtained is slightly less compared to the case of MgO \cite{DDFS} where 79 to 86 \% were recovered. One of the reasons for this difference is that we used a CPP in the case of MgO which covers nearly 100\% of the core-valence correlation contributions in Mg, while the explicit treatment of that correlation piece for Ca was less exhaustive. Another reason is that on the Hartree-Fock level $f$-functions for Ca (which are not yet implemented in CRYSTAL) would probably increase the cohesive energy and lower the 'experimental' correlation contribution. Finally, as in the case of MgO, a significant part of the missing correlation energy should be due to basis set errors for the O atom. -- The total cohesive energy recovered in our calculations is in the range of between 91 and 93 \% of the experimental value. Our results are compared in Table 5 to those from density functional calculations. We choose the results from \cite{DRFASH} where a correlation-only functional was used and 0.078 to 0.097 H of the correlation contribution to the cohesive energy were obtained, depending on the specific correlation functional used. \section{Lattice constants} At the Hartree-Fock level, the lattice constant is in good agreement with experiment for MgO \cite{CausaMgO1,DRFASH,CausaZupan,McCarthy,Catti}, whereas there is a deviation of 0.05 \AA \mbox{ } in the case of CaO \cite{MackrodtCaO}. It is interesting, therefore, to study the influence of correlation effects on lattice constants. In Tables 6 and 7, we give the necessary increments for MgO and CaO, respectively. We find two main effects of correlations. On the one hand, the van der Waals-interaction leads to a reduction of the lattice spacing since the attractive interaction is of the form $-\frac{1}{r^6}$ and obviously stronger at shorter distance. On the other hand, we find that the intra-ionic correlation of the $\rm O^{2-}$-ion forces a larger constant. This can be understood from the argument that excited configurations are lower in energy and mix more strongly with the ground-state determinant if the $\rm O^{2-}$ is less compressed as explained in section 3.2. Adding up all these contributions (cf.\ Table 8), they are found to nearly cancel in the case of MgO and to lead to a reduction of only 0.01 \AA \mbox{ }. For obtaining this result, we applied a linear fit to the correlation energy and superimposed it on the HF potential curve of Refs. \cite{Catti,MackrodtCaO}. We checked the validity of the linear approximation by calculating selected increments at other lattice constants. In the case of CaO, the van der Waals-interaction is more important and the lattice constant is reduced to 4.81 \AA \mbox{ } which is in nice agreement with the experimental value. The lattice constants seem to be in better agreement with the experimental values than those calculated from density functional theory for MgO \cite{CausaZupan,McCarthy} and CaO \cite{CausaZupan}, where deviations of $\pm$ 2 \% are found. This is similar to earlier findings for semiconductors \cite{Beate}. \section{Conclusion} We determined the correlation contribution to the cohesive energy of CaO using an expansion into local increments recently applied to MgO. Making use of quantum-chemical {\em ab-initio} configuration-interaction calculations for evaluating individual increments, we obtain $\sim$ 80 \% of the expected value. The missing energy is probably mainly due to the lack of $g$ and higher polarization functions in our one-particle basis set. The computed lattice constants show deviations of less than 1\% from the experimental values. We found two correlation effects on the lattice constants: the inter-atomic van der Waals-force leads to a reduction, whereas intra-atomic correlations of the $\rm O^{2-}$ ions lead to an increase of the lattice constant. The main difference between CaO and MgO is the reduced importance of the inter-atomic O-O correlations in CaO (due to the larger lattice constant) and the higher importance of the Ca-O correlations (due to the higher polarizability of Ca$^{2+}$). Compared to DFT, the numerical effort of our scheme is significantly higher. However, we feel that the advantage of the present approach is the high quality and stability of the results both for atoms, ions as well as for solids. Another advantage is the possibility of a systematic improvement by using larger basis sets. We think that the method of local increments is capable now of being routinely applied to ionic systems, and a systematic study on alkali halides is underway. An extension to open-shell systems such as NiO is also a project currently under investigation. \section*{Acknowledgments} We would like to thank Prof.\ P.\ Fulde for supporting this work and Prof.\ W.\ C.\ Nieuwpoort (Groningen) for interesting suggestions. We are grateful to Prof.\ H.-J.\ Werner (Stuttgart) for providing the program package MOLPRO.
proofpile-arXiv_065-498
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }
\section{INTRODUCTION} The discovery of at least five new $\gamma$-ray pulsars by the Compton Gamma-Ray Observatory (CGRO) and ROSAT has re-ignited theoretical work on the physical processes and modeling of high-energy radiation from pulsars. Including the previously known $\gamma$-ray pulsars, Crab (Nolan et al. 1993) and Vela (Kanbach et al. 1994), the recent detection of pulsed $\gamma$-rays from PSR B1706-44 (Thompson et al. 1992), Geminga (Halpern \& Holt 1992, Bertsch et al. 1992), PSR1509-58 (Wilson et al. 1993), PSR B1055-52 (Fierro et al. 1993), PSR B1951+32 (Ramanamurthy et al. 1995) and possibly PSR0656+14 (Ramanamurthy et al. 1996) bring the total to at least seven. For the first time, it is possible to look for the similarities and patterns in the $\gamma$-ray emission characteristics that may reveal clues to the origin of this emission. Examples are a preponderance of double pulses, a $\gamma$-ray luminosity vs. polar cap current correlation, a spectral hardness vs. characteristic age correlation, and spectral cutoffs above a few GeV (e.g. Thompson 1996). PSR1509-58 stands out among the known $\gamma$-ray pulsars as having both an unusually low spectral cutoff energy (somewhere between 2 and 30 MeV) and the highest inferred surface magnetic field ($3 \times 10^{13}$ Gauss). It has been detected by the CGRO instruments operating only in the lowest energy bands, BATSE (Wilson et al. 1993) and OSSE (Matz et al. 1994), with the higher energy instruments, COMPTEL (Bennett et al. 1993) and EGRET (Nel et al. 1996), giving upper limits that require a cutoff or turnover between 2 and 30 MeV. There is no evidence for pulsed TeV emission (Nel et al. 1992). There are currently two types of models for $\gamma$-ray pulsars being investigated in detail. Polar cap models assume that particles are accelerated along open field lines near the neutron star by strong parallel electric fields (e.g. Arons 1983). The primary particles induce electromagnetic cascades through the creation of electron-positron pairs by either curvature radiation (Daugherty \& Harding 1982, 1994, 1996) or inverse-Compton radiation (Sturner \& Dermer 1994) $\gamma$-rays. Outer gap models assume that the primary particles are accelerated along open field lines in the outer magnetosphere, near the null charge surface, where the corotation charge changes sign, and where strong electric fields may develop (Cheng, Ho \& Ruderman 1986a,b; Chiang \& Romani 1992; Romani \& Yadigaroglu 1995). Since the magnetic fields in the outer gaps are too low to sustain one-photon pair production cascades, these models must rely on photon-photon pair production of $\gamma$-rays, interacting with either non-thermal X-rays from the gap or thermal X-rays from the neutron star surface, to initiate pair cascades. Magnetic one-photon pair production, $\gamma\to e^+e^-$, has so far been the only photon attenuation mechanism assumed to operate in polar cap cascade models. Another attenuation mechanism, photon splitting \teq{\gamma\to\gamma\gamma}, will also operate in the high-field regions near pulsar polar caps but has not yet been included in polar cap model calculations. The rate of photon splitting increases rapidly with increasing field strength (Adler 1971), so that it may even be the dominant attenuation process in the highest field pulsars. There are several potentially important consequences of photon splitting for $\gamma$-ray pulsar models. Since photon splitting has no threshold, it can attenuate photons below the threshold for pair production, $\varepsilon = 2/\sin\theta_{\rm kB}$, and can thus produce cutoffs in the spectrum at lower energies. Here $\theta_{\rm kB}$ is the angle between the photon momentum and the magnetic field vectors, and $\varepsilon$ is (hereafter) expressed in units of $m_ec^2$. When the splitting rate becomes large enough, splitting can take place during a photon's propagation through the neutron star magnetosphere \it before \rm the pair production threshold is crossed (i.e. before an angle \teq{\sim 2/\varepsilon} to the field is achieved). Consequently, the production of secondary electrons and positrons in pair cascades will be suppressed. Instead of {\it pair} cascades, one could have {\it splitting} cascades, where the high-energy photons split repeatedly until they escape the magnetosphere. The potential importance of photon splitting in neutron star applications was suggested by Adler (1971), Mitrofanov \it et al. \rm (1986) and Baring (1988). Its attenuation and reprocessing properties have been explored in the contexts of annihilation line suppression in gamma-ray pulsars (Baring 1993), and spectral formation of gamma-ray bursts from neutron stars (Baring 1991). Photon splitting cascades have also been investigated in models of soft $\gamma$-ray repeaters, where they will soften the photon spectrum very efficiently with no production of pairs (Baring 1995, Baring \& Harding 1995a, Harding \& Baring 1996, Chang et al. 1996). In this paper we examine the importance of photon splitting in $\gamma$-ray pulsar polar cap models (it presumably will not operate in the low fields of outer gap models). Following a brief discussion of the physics of photon splitting in Section~2, we present calculations of the splitting attenuation lengths and escape energies in the dipole magnetic field of a neutron star. A preliminary study (Harding, Baring \& Gonthier 1996) has shown that splitting will be the primary mode of attenuation of $\gamma$-rays emitted parallel to a magnetic field \teq{B \gtrsim 0.3 B_{\rm cr} = 1.3\times 10^{13}} Gauss. We then present, in Section~3, photon splitting cascade models for two cases: (1) when only one mode of splitting ($\perp \rightarrow \parallel\parallel$) allowed by the kinematic selection rules (Adler 1971, Shabad 1975) operates, suppressing splitting of photons of parallel polarization (so that they can only pair produce), but still permitting photons of perpendicular polarization to either split once or produce pairs, and (2) when the three splitting modes allowed by CP (charge-parity) invariance operate, producing mode switching and a predominantly photon splitting cascade. In Section~4, model cascade spectra are compared to the observed spectrum of PSR1509-58 to determine the range of magnetic colatitude emission points (if any) that can produce a spectral cutoff consistent with the data. These spectra have cutoff energies that are decreasing functions of the magnetic colatitude. It is found that a reasonably broad range of polar cap sizes will accommodate the data, and that strong polarization signatures appear in the spectra due to the action of photon splitting. \section{PHOTON SPLITTING AND PAIR CREATION ATTENUATION} The basic features of magnetic photon splitting \teq{\gamma\to\gamma\gamma} and the more familiar process of single-photon pair creation \teq{\gamma\to e^+e^-} are outlined in the next two subsections before investigating their role as photon attenuation mechanisms in pulsar magnetospheres. Note that throughout this paper, energies will be rendered dimensionless, for simplicity, using the scaling factor \teq{m_ec^2}. Magnetic fields will also often be scaled by the critical field \teq{B_{\rm cr}}; this quantity will be denoted by a prime: \teq{B'=B/B_{\rm cr}}. \subsection{Photon Splitting Rates} The splitting of photons in two in the presence of a strong magnetic field is an exotic and comparatively recent prediction of quantum electrodynamics (QED), with the first correct calculations of the reaction rate being performed in the early '70s (Bialynicka-Birula and Bialynicki-Birula 1970; Adler \it et al. \rm 1970; Adler 1971). Its relative obscurity to date (compared, for example, with magnetic pair creation) in the astrophysical community stems partly from the mathematical complexity inherent in the computation of the rate. Splitting is a third-order QED process with a triangular Feynman diagram. Hence, though splitting is kinematically possible, when \teq{B=0} it is forbidden by a charge conjugation symmetry of QED known as Furry's theorem (e.g. see Jauch and Rohrlich, 1980), which states that ring diagrams that have an odd number of vertices with only external photon lines generate interaction matrix elements that are identically zero. This symmetry is broken by the presence of an external field. The splitting of photons is therefore a purely quantum effect, and has appreciable reaction rates only when the magnetic field is at least a significant fraction of the quantum critical field \teq{B_{\rm cr} = m_e^2c^3/(e\hbar )=4.413\times 10^{13}} Gauss. Splitting into more than two photons is prohibited in the limit of zero dispersion because of the lack of available quantum phase space (Minguzzi 1961). The reaction rate for splitting is immensely complicated by dispersive effects (e.g. Adler 1971; Stoneham 1979) caused by the deviation of the refractive index from unity in the strong field. Consequently, manageable expressions for the rate of splitting are only possible in the limit of zero dispersion, and are still then complicated triple integrations (see Stoneham 1979, and also Ba\u{\i}er, Mil'shte\u{\i}n, and Sha\u{\i}sultanov 1986 for electric field splitting) due to the presence of magnetic electron propagators in the matrix element. Hence simple expressions for the rate of splitting of a photon of energy \teq{\omega} in a field \teq{B} were first obtained by Bialynicka-Birula and Bialynicki-Birula (1970), Adler \it et al. \rm (1970) and Adler (1971) in the low-energy, non-dispersive limit: \teq{\omega B/B_{\rm cr}\lesssim 1}. The total rate in this limit, averaged over photon polarizations (Papanyan and Ritus, 1972), is expressible in terms of an attenuation coefficient \begin{equation} T_{\rm sp} (\omega )\;\approx\; {{\alpha^3}\over{10\pi^2}}\, \dover{1}{\lambda\llap {--}}\, {\biggl({{19}\over{315}}\biggr) }^2\, B'^6\, {\cal C}(B')\, \omega^5\,\sin^6\theta_{\rm kB} \quad , \label{eq:splitotrate} \end{equation} where \teq{\alpha =e^2/\hbar c\approx 1/137} is the fine structure constant, \teq{\lambda\llap {--} =\hbar /(m_ec)} is the Compton wavelength of the electron, and \teq{\theta_{\rm kB}} is the angle between the photon momentum and the magnetic field vectors. Here \teq{{\cal C}(B')} is a strong-field modification factor (derivable, for example, from Eq.~41 of Stoneham, 1979: see Eq.~[\ref{eq:splitratecorr}] below) that approximates unity when \teq{B\ll B_{\rm cr}} and scales as \teq{B^{-6}} for \teq{B\gg B_{\rm cr}}. The corresponding differential spectral rate for the splitting of photons of energy \teq{\omega} (with \teq{\omega\ll 1}) into photons of energies \teq{\omega'} and \teq{\omega -\omega'} is \begin{equation} T_{\rm sp}(\omega ,\omega')\;\approx\; 30\,\dover{\omega'^2 (\omega -\omega' )^2}{\omega^5}\, T_{\rm sp}(\omega )\quad . \label{eq:splitdiffrate} \end{equation} Equations~(\ref{eq:splitotrate}) and~(\ref{eq:splitdiffrate}) are valid (Baring, 1991) when \teq{\omega B' \sin\theta_{\rm kB}\lesssim 1}, which for pulsar fields and \teq{\omega \sin\theta_{\rm kB} \lesssim 2}, generally corresponds to the regime of weak vacuum dispersion. Reducing \teq{\theta_{\rm kB}} or \teq{B} dramatically increases the photon energy required for splitting to operate in a neutron star environment. The produced photons emerge at an angle \teq{\theta_{\rm kB}} to the field since splitting is a collinear process in the low-dispersion limit. Adler (1971) observed that in the low-energy limit, the splitting rate was strongly dependent on the polarization states of the initial and final photons; this feature prompted the suggestion by Adler et al. (1970) and Usov and Shabad (1983) that photon splitting should be a powerful polarizing mechanism in pulsars. The polarization-dependent rates can be taken from Eq.~(23) of Adler (1971), which can be related to equations~(\ref{eq:splitotrate}) or~(\ref{eq:splitdiffrate}) via \begin{equation} T^{\rm sp}_{\perp\to\parallel\parallel}\; =\; \dover{1}{2}\, T^{\rm sp}_{\parallel\to\perp\parallel}\; =\; \biggl( \dover{{\cal M}_1^2}{{\cal M}_2^2} {\biggr)}^2\, T^{\rm sp}_{\perp\to\perp\perp}\; =\; \dover{2 {\cal M}_1^2\, T_{\rm sp} }{ 3{\cal M}_1^2 + {\cal M}_2^2 } \quad , \label{eq:splitpolrate} \end{equation} where the scattering amplitude coefficients \begin{eqnarray} {\cal M}_1 & = &\dover{1}{B'^4}\int^{\infty}_{0} \dover{ds}{s}\, e^{-s/B'}\, \Biggl\{ \biggl(-\dover{3}{4s}+\dover{s}{6}\biggr)\,\dover{\cosh s}{\sinh s} +\dover{3+2s^2}{12\sinh^2s}+\dover{s\cosh s}{2\sinh^3s}\Biggr\}\nonumber\\ {\cal M}_2 & = &\dover{1}{B'^4}\int^{\infty}_{0} \dover{ds}{s}\, e^{-s/B'}\, \Biggl\{ \dover{3}{4s}\,\dover{\cosh s}{\sinh s} + \dover{3-4s^2}{4\sinh^2s} - \dover{3s^2}{2\sinh^4s}\Biggr\} \label{eq:splitcoeff} \end{eqnarray} are given in Adler (1971) and Eq.~41 of Stoneham (1979). In the limit of \teq{B\ll B_{\rm cr}}, \teq{{\cal M}_1\approx 26/315} and \teq{{\cal M}_2\approx 48/315}, while in the limit of \teq{B\gg B_{\rm cr}}, equation~(\ref{eq:splitcoeff}) produces \teq{{\cal M}_1\approx 1/(6B'^3)} and \teq{{\cal M}_2\approx 1/(3B'^4)}. The factor of two in the numerator of the right hand side of equation~(\ref{eq:splitpolrate}) accounts for the duplicity of photons produced in splitting. The photon polarization labelling convention of Stoneham (1979) is adopted here (this standard form was not used by Adler, 1971): the label \teq{\parallel} refers to the state with the photon's \it electric \rm field vector parallel to the plane containing the magnetic field and the photon's momentum vector, while \teq{\perp} denotes the photon's electric field vector being normal to this plane. Summing over the polarization modes yields the relationship for the strong-field modification factor in equation~(\ref{eq:splitotrate}): \begin{equation} {\cal C}(B')\; =\;\dover{1}{12}\, {\biggl({{315}\over{19}}\biggr) }^2\, \Bigl( 3{\cal M}_1^2 + {\cal M}_2^2 \Bigr) \quad . \label{eq:splitratecorr} \end{equation} Note that, in the absence of vacuum dispersion, the splitting modes \teq{\perp\to\perp\parallel}, \teq{\parallel\to\perp\perp} and \teq{\parallel\to\parallel\parallel} are forbidden by arguments of CP (charge-parity) invariance in QED (Adler 1971); dispersive effects admit the possibility of non-collinear photon splitting so that there is a small but non-zero probability for the \teq{\perp\to\perp\parallel} channel. Equations~(\ref{eq:splitotrate})--(\ref{eq:splitratecorr}) define the rates to be used in the analyses of this paper, and are valid for \teq{\omega B'\sin\theta_{\rm kB}\ll 1}. The triple integral expressions that Stoneham (1979) derives are valid (below pair creation threshold) for a complete range (i.e. 0 to \teq{\infty}) of the expansion parameter \teq{\omega B'\sin\theta_{\rm kB}}, but are not presently in a computational form suitable for use here. Work is in progress to address this deficiency (Baring \& Harding 1996), and preliminary results indicate that equations~(\ref{eq:splitotrate})--(\ref{eq:splitratecorr}) approximate Stoneham's (1979) formulae to better than two percent for \teq{\omega B'\sin\theta_{\rm kB}\leq 0.2}, and differs by at most a factor of around 2.5 for \teq{\omega B'\sin\theta_{\rm kB} \sim 1.5}, the value relevant to the calculations of this paper; the splitting rate given by Stoneham's formulae initially increase above the low energy limits as \teq{\omega B'\sin\theta_{\rm kB}} increases. Recently there has appeared a new result on the rates of photon splitting. Mentzel, Berg \& Wunner (1994) presented an S-matrix calculation of the rates for the three polarization modes permitted by CP invariance that are considered here. While their formal development is comparable to an earlier S-matrix formulation of splitting in Melrose \& Parle (1983a,b), their presentation of numerical results appeared to be in violent disagreement (see also their astrophysical presentation in Wunner, Sang \& Berg 1995) with the splitting results obtained via the Schwinger proper-time technique by Adler (1971) and Stoneham (1979) that comprise equations~(\ref{eq:splitotrate})--(\ref{eq:splitratecorr}) here. These results have now been retracted, the disagreement being due to a sign error in their numerical code (Wilke \& Wunner 1996). The revised results are in much better agreement with the rates computed by Adler (1971). However, the revised numerical splitting rates of Wilke \& Wunner (1996) still differ by as much as a factor of 3 from Baring \& Harding's (1996) computations of Stoneham's (1979) general formulae. Ba\u{\i}er, Mil'shte\u{\i}n, \& Sha\u{\i}sultanov (1996) generate numerical results from their earlier alternative proper-time calculation (Ba\u{\i}er, Mil'shte\u{\i}n, \& Sha\u{\i}sultanov 1986) that are in accord with Stoneham's and Adler's (1971) results and also with those of Baring \& Harding (1996). The numerical computation of the S-matrix formalism is a formidable task. The proper-time analysis, though difficult, is more amenable, and has been reproduced in the limit of \teq{B\ll B_{\rm cr}} by numerous authors. As the S-matrix and proper-time techniques should produce equivalent results, and indeed have done so demonstrably in the case of pair production (see DH83 and Tsai \& Erber 1974), we choose to use the amenable proper-time results outlined above in the calculations of this paper. The above results ignore the fact that the magnetized vacuum is dispersive and birefringent, so that the phase velocity of light is less than \teq{c} and depends on the photon polarization. Dispersion can therefore alter the kinematics of QED processes such as splitting (Adler 1971), and further dramatically complicates the formalism for the rates (Stoneham 1979). Extensive discussions of dispersion in a magnetized vacuum are presented by Adler (1971) and Shabad (1975); considerations of plasma dispersion are not relevant to the problem of gamma-ray emission from pulsars because they become significant only for densities in excess of around \teq{10^{27}}cm$^{-3}$, which are only attained at the stellar surface. Adler (1971) showed that in the limit of \it weak \rm vacuum dispersion (roughly delineated by $B'\sin\theta_{\rm kB} \lesssim 1$), where the refractive indices for the polarization states are {\it very} close to unity, energy and momentum could be simultaneously conserved only for the splitting mode \teq{\perp\to\parallel\parallel} (of the modes permitted by CP invariance) below pair production threshold. This kinematic selection rule was demonstrated for subcritical fields, where the dispersion is very weak, a regime that generally applies to gamma-ray pulsar scenarios. Therefore, it is probable that only the one mode (\teq{\perp\to\parallel\parallel}) of splitting operates in gamma-ray pulsars. This result may be modified by subtle effects such as those incurred by field non-uniformity. We adopt a dual scenario in this paper for the sake of completeness: one in which all CP-permitted modes of splitting operate, and one in which Adler's kinematic selection rules are imposed. Note that in magnetar models of soft gamma repeaters (e.g Baring 1995, Harding and Baring 1996), where supercritical fields are employed, moderate vacuum dispersion arises. In such a regime, it is not clear whether Adler's selection rules still endure, since his analysis implicitly uses weak dispersion limits of linear vacuum polarization results (e.g. see Shabad 1975), and omits higher order contributions (e.g. see Melrose and Parle 1983a,b) to the vacuum polarization (for example, those that couple to photon absorption via splitting) that become prominent in supercritical fields. Furthermore, plasma dispersion effects may be quite pertinent to soft gamma repeater models (e.g. Bulik and Miller 1996), rendering them distinctly different from pulsar scenarios. \subsection{Pair Production Rate} One-photon pair production is a first-order QED process that is quite familiar to pulsar theorists. It is forbidden in field-free regions due to the imposition of four-momentum conservation, but takes place in an external magnetic field, which can absorb momentum perpendicular to \bf B\rm . The rate (Toll 1952, Klepikov 1954) increases rapidly with increasing photon energy and transverse magnetic field strength, becoming significant for $\gamma$-rays above the threshold, $\omega = 2/\sin\theta_{\rm kB}$, and for fields approaching $B_{\rm cr}$. When the photon energy is near threshold, there may be only a few kinematically available pair states, and the rate will be resonant at each pair state threshold, producing a sawtooth structure (Daugherty \& Harding 1983, hereafter DH83). For photon energies well above threshold, the number of pair states becomes large, allowing the use of a more convenient asymptotic expression for the polarization dependent attenuation coefficient (Klepikov 1954, Tsai \& Erber 1974): \begin{equation} T^{\rm pp}_{\parallel,\perp} = {1\over 2}{\alpha\over \lambda\llap {--}} B' \sin\theta_{\rm kB}\Lambda_{\parallel,\perp}(\chi), \label{eq:ppasymp} \end{equation} \begin{equation} \Lambda_{\parallel, \perp}(\chi) \approx \left\{ \begin{array}{lr} (0.31, 0.15)\, \exp \mbox{\Large $(-{4\over 3\chi})$} & \chi \ll 1 \\ \\ (0.72, 0.48) \, \chi^{-1/3} & \chi \gg 1 \end{array} \right. \label{eq:ppratlim} \end{equation} where $\chi \equiv (\omega/2)B'\sin\theta_{\rm kB}$. In polar cap pulsar models (e.g. Sturrock 1971, Ruderman and Sutherland 1975), high energy radiation is usually emitted at very small angles to the magnetic field, well below pair threshold (see Harding 1995, for review). The $\gamma$-ray photons will convert into pairs only after they have traveled a distance $s$ comparable to the field line radius of curvature $\rho$, so that $\sin\theta_{\rm kB} \sim s/\rho$. From the above expression, the pair production rate will be vanishingly small until the argument of the exponential approaches unity, i.e. when $\omega B'\sin\theta_{\rm kB} \gtrsim 0.2$. Consequently, pair production will occur well above threshold when $B \ll 0.1 B_{\rm cr}$ and the asymptotic expression will be valid. However when $B \gtrsim 0.1 B_{\rm cr}$, pair production will occur at or near threshold, where the asymptotic expression has been shown to fall orders of magnitude below the exact rate (DH83). In the present calculation, we approximate the near-threshold reduction in the asymptotic pair production attenuation coefficient by making the substitution, $\chi \rightarrow \chi/F$, where $F = 1 + 0.42(\omega\sin\theta_{\rm kB}/2)^{-2.7}$ in equation~(\ref{eq:ppratlim}) (DH83). Baring (1988) has derived an analytic expression for the one-photon pair production rate near threshold which gives a result that agrees numerically with the approximation of DH83. Yet even the near-threshold correction to the asymptotic rate becomes poor when $B \gg 0.1 B_{\rm cr}$ and the photons with parallel and perpendicular polarization produce pairs only (DH83) in the ground (0,0) and first excited (0,1) and (1,0) states respectively. Here \teq{(j,k)} denotes the Landau level quantum numbers of the produced pairs. Therefore when the local $B>0.1 B_{\rm cr}$, instead of the asymptotic form in equation~(\ref{eq:ppratlim}), we use the exact, polarization-dependent, pair production attenuation coefficient (DH83), including only the (0,0) pair state for $\parallel$ polarization: \begin{equation} T^{\rm pp}_{\parallel} = \dover{\alpha\sin\theta_{\rm kB}}{\lambda\llap {--} \xi |p_{\hbox{\sevenrm 00}}|}\,\exp(-\xi), \quad \omega \ge 2/\sin\theta_{\rm kB}\quad , \label{eq:tpppar} \end{equation} and only the sum of the (0,1) and (1,0) states for $\perp$ polarization: \begin{equation} T^{\rm pp}_{\perp} = \dover{\alpha\sin\theta_{\rm kB}}{\lambda\llap {--}\xi |p_{\hbox{\sevenrm 01}}|}\, (E_0 E_1 + 1 + p_{01}^2)\,\exp(-\xi), \quad \omega \ge \dover{1+(1 + 2B')^{1/2}}{\sin\theta_{\rm kB}} \label{eq:tppperp} \end{equation} where \begin{displaymath} E_0 = (1 + p_{01}^2)^{1/2}\quad , \quad E_1 = (1 + p_{01}^2 + 2B')^{1/2} \end{displaymath} for \begin{displaymath} |p_{jk}| = \left[ \dover{\omega^2}{4} \sin^2\theta_{\rm kB} - 1 - (j+k)B' + \left( \dover{(j-k)B'}{\omega\sin\theta_{\rm kB}} \right)^2\right]^{1/2} \end{displaymath} and \begin{equation} \xi = \dover{\omega^2}{2B'}\sin^2\theta_{\rm kB}\quad . \label{eq:xi} \end{equation} Actually, both the asymptotic and exact mean-free paths ($1/T_{\rm pp}$) are so small in fields where photons pair produce at threshold that it is, in fact, not important which rate is used at very high field strengths (i.e. $B' \gtrsim 1$). The pair production rate in this regime thus behaves like a wall at threshold and photons will pair produce as soon as they satisfy the kinematic restrictions on \teq{\omega} given in equations~(\ref{eq:tpppar}) and~(\ref{eq:tppperp}). The creation of bound pairs rather than free pairs is possible in fields $B' \gtrsim 0.1$ (Usov \& Melrose 1995), but this should not affect the present calculation since we do not model the full pair cascade. \subsection{Attenuation Lengths} To assess the relative importance of photon splitting compared to pair production through a dipole magnetic field, we compute the attenuation length \teq{L}, defined to be the path length over which the optical depth is unity: \begin{equation} \tau(\theta, \varepsilon)\; =\;\int_0^L T(\theta_{\rm kB},\, \omega )\, ds\; =\; 1\quad , \label{eq:tau} \end{equation} where \teq{ds} is the pathlength differential along the photon momentum vector \teq{{\bf k}}and $T$ is the attenuation coefficient for either splitting, $T_{\rm sp}$, or pair production, $T_{\rm pp}$. In this paper, attenuation lengths are computed as averages over polarizations of the initial photon and, for splitting, sums over the final polarization states. Here \teq{\theta_{\rm kB}} and the photon energy \teq{k^{\hat o} = \omega} are functions of the position (e.g. see equation~[\ref{eq:4momgr}]), specifically measured in the local inertial frame, while \teq{\theta} is the colatitude of emission and \teq{\varepsilon} is the photon energy to an observer at infinity; our treatment of curved spacetime is discussed immediately below. In regions where the path length is much shorter than both the scale length of the field strength variation or the radius of curvature of the field, \teq{L} reduces to the inverse of the attenuation coefficient. In the calculation of the splitting attenuation lengths, all three CP-permitted modes are assumed to operate. The attenuation length behavior of the individual modes are similar. \placefigure{fig:geometry} We assume that test photons are emitted at the neutron star surface and propagate outward, initially parallel or at a specified angle, $\theta_{\rm kB,0}$ to the dipole magnetic field (see Fig.~\ref{fig:geometry} for a depiction of the geometry). Photon emission in polar cap models of gamma-ray pulsars can occur above the stellar surface (but see the discussion in Section~5), which would generate attenuation lengths somewhat longer than those determined here, due to the \teq{r^{-3}} decay of the field. A surface origin of the photons is chosen in this paper to provide a simple and concise presentation of the attenuation properties. We have included the general relativistic effects of curved spacetime in a Schwarzschild metric, following the treatment of Gonthier \& Harding (1994, GH94) who studied the effects of general relativity on photon attenuation via magnetic pair production. GH94 included the curved spacetime photon trajectories, the magnetic dipole field in a Schwarzschild metric and the gravitational redshift of the photon energy. One improvement we have made here to the treatment of GH94 is to explicitly keep track of the gravitational redshift of the photon energy as a function of distance from the neutron star (see Appendix for details). Our analysis is confined to the Schwarzschild metric because the dynamical timescales for gamma-ray pulsars are considerably shorter than their period (e.g. \teq{P=0.15}sec. for PSR1509-58), so that rotation effects in the Kerr metric can be neglected. We have taken a neutron star mass, $M = 1.4\,M_\odot$ and radius, $R = 10^6$ cm in these calculations. Fig.~\ref{fig:attenl} illustrates how the attenuation lengths for photon splitting and pair production vary with energy for different magnetic colatitudes of the emission point, for surface fields of \teq{B_0=0.1B_{\rm cr}}, and \teq{B_0=0.7B_{\rm cr}}. A field of $B_0 = 0.7B_{\rm cr}$ is the value of the polar surface field derived from the magnetic dipole spin-down energy loss (Shapiro \& Teukolsky 1983), using the measured $P$ and $\dot P$ for PSR1509-58. As noted by Usov \& Melrose (1995), this is exactly twice the value of the surface field given by formulae in other sources (Manchester \& Taylor 1977, Michel 1991), which assume (inaccurately) that the dipole magnetic moment $\mu = B_0\,R^3$ rather than $\mu = B_0\,R^3/2$ for a uniformly magnetized sphere of radius $R$. The other $\gamma$-ray pulsars have surface field strengths in the range $1 - 9\times 10^{12}$ G, or $0.02 - 0.2\,B_{\rm cr}$ (the Crab and Vela pulsars have fields around $0.2\,B_{\rm cr}$). Note that the attenuation lengths in Fig.~\ref{fig:attenl} are for unpolarized radiation; the curves for $\parallel$ and $\perp$ polarization states look very similar. \placefigure{fig:attenl} The curves in Fig.~\ref{fig:attenl} have a power-law behavior at high energies, i.e. for attenuation lengths much less than $10^{6}$ cm, where the dipole field is almost uniform in direction and of roughly constant strength. They also exhibit sharp increases at the low energy end, where photons begin to escape the magnetosphere without attenuation. We may estimate the behavior of the power-law portions of the attenuation length curves in Fig~\ref{fig:attenl} as follows. Since the photons are assumed to initially propagate parallel to the field, the field curvature will give propagation oblique to the field only after significant distances are traversed, so that the obliquity of the photon to the field scales, to first order, as the distance travelled, \teq{\sin\theta_{\rm kB}\propto s}. Inserting this in equation~(\ref{eq:splitotrate}) gives a photon splitting attenuation coefficient \teq{\propto s^6} i.e. an optical depth \teq{\propto \varepsilon^5 s^7}, since \teq{T_{\rm sp}\propto\varepsilon^5}. Inversion then indicates that the attenuation length should vary as \teq{L \propto \varepsilon^{-5/7}}: this is borne out in Fig.~\ref{fig:attenl}. For $B_0 \gtrsim 0.1B_{\rm cr}$, pair production occurs as soon as the threshold $\varepsilon_{\rm th} = 2/\sin\theta_{\rm kB}$ is crossed (cf. Section 2.2) during the photon propagation in the magnetosphere. Essentially, due to the enormous creation rate immediately above the threshold, this energy serves as an impenetrable ``wall'' to the photon. Again, since \teq{\sin\theta_{\rm kB} \propto s} in the early stages of propagation, the pair production attenuation length should scale as \teq{L\propto 2/\varepsilon}. These proportionalities hold in both curved and flat spacetime since general relativistic effects distort spacetime in a smooth and differentiable manner (see the Appendix). However, the attenuation lengths computed in the Schwarzschild metric are about a factor of 1.5 lower than those computed in flat spacetime (Baring \& Harding 1995b). The photon splitting attenuation coefficient we have used is strictly valid only below pair threshold. Hence, the attenuation lengths for splitting depicted in Fig.~\ref{fig:attenl} can be regarded as only being symbolic when they exceed those for pair production, since then pair threshold is reached before splitting occurs. No technically amenable general expressions for the rate of splitting above pair threshold exist in the physics literature. But the vicinity of parameter space just below pair threshold is the regime of importance for $\gamma$-ray pulsar models, where the emitted photons propagate until they either split or they reach pair threshold, in which case they pair produce. The attenuation length curves near the crossover points in Fig.~\ref{fig:attenl} for $B_0 = 0.7B_{\rm cr}$ will require inclusion of high energy corrections to the attenuation coefficient (Stoneham 1979) that arise as the $\gamma\to e^+e^-$ threshold is approached. Currently work is in progress to compute these modifications (Baring and Harding 1996, in preparation), and preliminary results indicate that the rate in equation~(\ref{eq:splitotrate}) is quite accurate for \teq{B\lesssim 0.2B_{\rm cr}}, but increases by factors of at most a few for \teq{B=0.7B_{\rm cr}} and \teq{\omega =2}, as mentioned in Section~2.1 above. \subsection{Escape Energies} The energy at which the attenuation length becomes infinite defines the \it escape energy\rm, below which the optical depth is always \teq{\ll 1}, and photons escape the magnetosphere; the existence of such an escape energy is a consequence of the \teq{r^{-3}} decay of the dipole field. Escape energies of unpolarized photons for both photon splitting and pair production are shown in Fig.~\ref{fig:escape} as a function of magnetic colatitude $\theta$ of the photon emission point for different values of magnetic field strength (see also Harding, Baring and Gonthier, 1996). The escape energies clearly decline with \teq{\theta} and are monotonically decreasing functions of \teq{B} for the range of fields shown. The divergences as \teq{\theta\to 0} are due to the divergence of the field line radius of curvature at the poles. There the maximum angle \teq{\theta_{\rm kB}} achieved before the field falls off and inhibits attenuation is proportional to the colatitude \teq{\theta}. For photon splitting, since the rate in equation~(\ref{eq:splitotrate}) is proportional to \teq{\omega^5\sin^6\theta_{\rm kB}}, and therefore also the attenuation length \teq{L}, it follows that the escape energy scales as \teq{\varepsilon_{\rm esc} \propto \theta^{-6/5}} near the poles (see also Fig.~\ref{fig:icescape}) as is determined by the condition \teq{L\sim R}. For pair production, the behaviour of the rate (and therefore \teq{L}) is dominated by the exponential form in equation~(\ref{eq:ppratlim}), which then quickly yields a dependence \teq{\varepsilon_{\rm esc} \propto \theta^{-1}} near the poles for \teq{B_0\lesssim 0.1B_{\rm cr}}. This behaviour extends to higher surface fields because production then is at threshold, which determines \teq{\varepsilon_{\rm esc}\sim 2/\theta_{\rm kB} \propto \theta^{-1}}. At high fields, $B_0 \gtrsim 0.3B_{\rm cr}$, there is a saturation of the photon splitting attenuation lengths and escape energies, due to the diminishing dependence of $B$ in the attenuation coefficient. Likewise, there is a saturation of the pair production escape energy at fields above which pair production occurs at threshold. The pair production escape energy curves are bounded below by the pair threshold $2/\sin\theta_{\rm kB}$ and merge for high $\theta$, at the pair rest mass limit, $\varepsilon = 2$, blueshifted by the factor $(1-2GM/Rc^2)^{-1/2} \sim 1.3$. Note that photon splitting can attenuate photons well below pair threshold. For low fields, pair production escape energies are below those for splitting, but in high fields, splitting escape energies are lower at all $\theta$. The escape energies are roughly equal for $B_0 \sim 0.3 B_{\rm cr}$. \placefigure{fig:escape} The effects of curved spacetime are quite significant when compared to the attenuation lengths and the escape energies obtained assuming flat spacetime. A comparison of the escape energies for splitting and pair production, computed in flat and curved spacetime, is shown in Fig.~\ref{fig:escurvflat}. The largest effects are due to the increase of the surface dipole field strength by roughly a factor of 1.4, and the correction for the gravitational redshift of the photon, which increases the photon energy by roughly a factor 1.2 in the local inertial frame at the neutron star surface compared to the energy measured by the observer in flat space (see the Appendix). The combination of these effects decreases the photon splitting escape energy by a factor of about 2 compared to flat spacetime. The decrease in escape energy for pair production is also a factor of about 2, except at the largest values of $\theta$ and $B'$, where the pair rest mass limit is reached (cf. Fig.~\ref{fig:escape}) The escape energy is then no longer dependent on field strength, and the ratio of the curved to flat space escape energy is just the redshift of the photon energy ($\sim 0.8$) from the conversion point. This is achieved in the upper right hand corner of the figure; photon splitting has no such strict limit. The ratios also become insensitive to \teq{\theta} near the poles since there the photons move almost radially, thus traveling along straight trajectories, and the curved-space correction to the field is not changing rapidly with colatitude. The curvature of the photon trajectory in a Schwarzschild metric does not affect the escape energies, to first order, except in the case of emission at large colatitudes, where the photon wavevector makes a large angle to the radial direction. \placefigure{fig:escurvflat} High energy emission from curvature radiation, inverse Compton or synchrotron by relativistic particles with Lorentz factor $\Gamma$ will not beam the photons precisely along the magnetic field, but within some angle $\sim 1/\Gamma$ to the field. Fig.~\ref{fig:icescape} illustrates the effect on the escape energies of a non-zero angle of emission of the photons, for the case where the photons are emitted at angles toward the dipole axis. We have chosen the angle $\theta_{\rm kB,0} = .01$ rad $(= 0.57^\circ$) because it would be the angle at which photons with $\epsilon \sim 100$ would be emitted through the cyclotron upscattering process, $\theta_{\rm kB,0} \simeq B'/\epsilon$ (Dermer 1990). For emission angles $\theta_{\rm kB,0} = 0$ in Fig.~\ref{fig:icescape}a, which plots \teq{\varepsilon_{\rm esc}} for photon splitting, $\varepsilon_{\rm esc} \propto B_0^{-6/5}$ for $B_0 \lesssim 0.3B_{\rm cr}$ and $\varepsilon_{\rm esc} \propto \theta^{-6/5}$, for $\theta \lesssim 20^\circ$, dependences that naturally follow from the form of equation~(\ref{eq:splitotrate}). Generally, the escape energy is insensitive to the emission angle for $\theta \gtrsim 10 \theta_{\rm kB,0}$. For small angles, the escape energy decreases and the $\theta_{\rm kB,0} = .57^\circ$ curves flatten below the $\theta_{\rm kB,0} = 0$ curves, converging as \teq{\theta\to 0} to an energy that is proportional to \teq{(B_0\sin\theta_{\rm kB,0})^{-6/5}} (see Eq.~[\ref{eq:ergs}]). This convergence is a consequence of the field along photon trajectories that originate near the pole being almost uniform and tilted at about angle \teq{\theta_{\rm kB,0}} to the photon path. In Fig.~\ref{fig:icescape}b, the same effect is seen for pair creation, but this time the ``saturation'' is at the redshifted threshold energy $2(1-2GM/Rc^2)^{1/2}/\sin\theta_{\rm kB,0}$, and is independent of \teq{B_0}. We note that this behaviour at low colatitudes was observed, in the case of pair creation in flat spacetime, by Chang, Chen and Ho (1996). \placefigure{fig:icescape} An obvious exception to this expected behaviour is seen in Fig.~\ref{fig:icescape}b for the $B' = 3.1$ curve, where the escape energy is actually \it larger \rm at small colatitudes (\teq{1^\circ\lesssim\theta\lesssim 10^\circ}) when the emission angle $\theta_{\rm kB,0}$ is increased. This counter-intuitive result can be understood with the aid of Fig.~\ref{fig:wsinthet}, which shows the increase in $\sin\theta_{\rm kB}$, and $\omega\sin\theta_{\rm kB}$, determined in the local inertial frame, with path length \teq{s} along the photon trajectory. Note that (i) the \teq{\theta_{kB,0}=0} curves increase in proportion to \teq{s} when \teq{s/R\ll 1}, as described in the Appendix, and (ii) the $\sin\theta_{\rm kB}$ curves increase logarithmically with $s/R$ when \teq{s/R} is not very small. In this large field, pair production occurs when the threshold $\omega\sin\theta_{\rm kB} = 2$ is crossed, at the same path length for both $\theta_{\rm kB,0} = 0$ and $\theta_{\rm kB,0} = 0.57^\circ$. The differences in the photon trajectories (which are almost radial) for these two cases are so small that \teq{s} effectively represents the same height above the stellar surface for both \teq{\theta_{\rm kB,0}}. Since \teq{\omega\sin\theta_{\rm kB}\approx 2} defines the pair creation ``wall'' for both photon paths, the only difference in escape energies is due to the factor of \teq{\sin\theta_{\rm kB}} at the front of the pair creation rates in equations~(\ref{eq:ppasymp})--(\ref{eq:tppperp}). Hence, at the point of pair creation, the value of $\sin\theta_{\rm kB}$ is smaller for the $\theta_{\rm kB,0} = 0.57^\circ$ case, and therefore the escape energy is larger. In flat spacetime, which is not depicted in Figs.~\ref{fig:icescape} or~\ref{fig:wsinthet}, the crossover point of the $\sin\theta_{\rm kB}$ curves occurs at the same $s/R$ value as pair threshold, so that the escape energies are the same at this colatitude for the two cases (this situation was also observed by Chang, Chen and Ho 1996). Note that as photon splitting does not have the same sudden onset as pair creation, it takes place over a range of path lengths, mostly around \teq{0.1\lesssim s/R\lesssim 2}. Over this range, \teq{\sin\theta_{\rm kB}} in Fig.~\ref{fig:wsinthet} is generally larger for the \teq{\theta_{\rm kB,0}=0.57^\circ} case so that the splitting escape energy is correspondingly shorter than for emission parallel to the field, as is evident in Fig.~\ref{fig:icescape}a. \placefigure{fig:wsinthet} \section{CASCADE SPECTRA} Here we describe briefly our Monte Carlo simulation of photon propagation and attenuation via splitting and pair creation in neutron star magnetospheres, together with results for single (Section 3.2) and multiple (Section 3.3) generations of photon splitting. \subsection{Monte Carlo Calculation} We model the spectrum of escaping photons from a cascade above a neutron star polar cap, including both photon splitting and pair production, by means of a Monte Carlo simulation. The free parameters specified at the start of the calculation are the magnetic colatitude $\theta$, the angles $\theta_k$ and $\phi_k$ (see Fig.~\ref{fig:geometry}), the spectrum, the height above the surface $z_0 = r-R$ of the photon emission, and the surface magnetic field strength $B_0$ (note that entities with subscripts `0' designate determination at the stellar surface). From these quantities, and assuming that $\phi = 0$ without loss of generality, we compute the four-vectors of the photon position and momentum that are carried through the computation. Injected photons are sampled from a power-law distribution, \begin{equation} N(\varepsilon) = N_0\varepsilon^{-\alpha}, \quad \varepsilon_{min} < \varepsilon < \varepsilon_{max} \label{eq:N} \end{equation} Polarization is chosen randomly to simulate unpolarized emission; this can be altered, as desired, for any postulated emission mechanism. The path of each input photon is traced through the magnetic field, in curved spacetime, accumulating the survival probabilities for splitting, $P_{\rm surv}^s$, and for pair production, $P_{\rm surv}^p$, independently: \begin{equation} P_{\rm surv}(s) = \exp\Bigl\{-\tau(s)\Bigr\} \end{equation} where \begin{equation} \tau(s) = \int_0^s T(\theta_{\rm kB}, \omega ) ds' \end{equation} is the optical depth along the path. These survival probabilities implicitly depend on the origin \teq{{\bf r}_0} of the photon and its energy \teq{\varepsilon} at infinity. In computing the attenuation lengths (Section 2.3), the photon was assumed to split when the survival probability reaches $1/e$, i.e. when equation~(\ref{eq:tau}) is satisfied. In the cascade simulation, the photon may either split or pair produce. The fate of each cascade photon is determined as follows: if the combined survival probability, $P_{\rm surv}^sP_{\rm surv}^p > \Re_1$, where $\Re_1$ is a random number between 0 and 1, chosen at the emission point, then the photon escapes; if not, then if the probability that the photon survives splitting but not pair production, $P_{\rm surv}^s(1-P_{\rm surv}^p)/(1-P_{\rm surv}^s P_{\rm surv}^p) > \Re_2$, where $\Re_2$ is a second random number, then the photon pair produces; otherwise, the photon splits. When the photon splits, the energy of one of the final photons is sampled from the distribution given in equation~(\ref{eq:splitdiffrate}) and their polarizations are chosen from the branching ratios given in equation~(\ref{eq:splitpolrate}). The energy of the second photon from the splitting is determined simply from energy conservation, since both final photons are assumed to be collinear in the direction of the parent photon. Each final photon is then followed in the same way as the injected photon, with a call to a recursive procedure that stores photon energies and positions through many generations of splitting. When the photon pair produces, the code does not follow the radiation from the pairs but simply returns to the previous cascade generation. For field strengths typical of gamma-ray pulsars, the pair radiation, most probably synchrotron or inverse Compton, will not contribute significantly at the energies near the escape energy for the cascades where all splitting modes operate. An exception to this may occur for supercritical surface fields, where synchrotron photons acquire most of the energy of their primary electrons. When all splitting modes operate, the number of pair production events is a small fraction of the number of splitting events for $B_0 = 0.7B_{\rm cr}$. The cascade photons are followed through many generations of splitting until all of the photons either escape or pair produce. The escaping photons are binned in energy and polarization. \subsection{Partial Splitting Cascade} For pulsar applications with subcritical fields, as discussed in Section 2.1, it is probable that the splitting modes allowed by CP invariance are further limited by kinematic selection rules to only the $\perp \rightarrow \parallel\parallel$ mode. This restriction may be confined to regimes of weak vacuum dispersion and may also depend on subtleties such as field non-uniformity. Such selection rules would effectively prevent splitting cascades since $\perp$ photons could split only into $\parallel$ photons which do not split. Here we compute the emergent spectra in this type of cascade, a partial cascade, where $\perp$ mode photons can either pair produce or split into $\parallel$ mode photons, while the $\parallel$ mode photons may only pair produce. There is a limit of two cascade generations: one splitting and one pair production. The input spectrum is a power-law (Eq.~[\ref{eq:N}]) with the parameters: $\varepsilon_{min} = 10^{-3}$, $\varepsilon_{max} = 100$ and $\alpha = 1.6$. The value of the index $\alpha$ is chosen to match the power-law fit of the OSSE spectrum of PSR1509-58 (Matz et al. 1994). The maximum energy of the input spectrum $\varepsilon_{max} = 100$ is chosen to fall above the 30 MeV maximum possible cutoff or turnover energy of the observed PSR1509-58 spectrum. For these runs, injection of 5 - 10 million photons are required to give adequate statistics. The number of pairs produced relative to photons in these partial splitting cascades is obviously higher than in the full cascades examined in the next section. Note that in more complete gamma-ray pulsar models that include the pair radiation, multiple generations of splitting might still be possible, being interspersed with generations of conventional synchrotron/pair cascading. \placefigure{fig:specpar} Figure~\ref{fig:specpar} shows partial splitting cascade spectra in each final polarization mode for photons injected parallel to the local magnetic field at different magnetic colatitudes. The spectra for the two polarization modes are cutoff at slightly different energies, reflecting the different escape energies for splitting, which cuts off the $\perp$ mode, and for pair production, which cuts off the $\parallel$ mode. There is a slight bump below the cutoff in the $\parallel$ mode spectrum, due to escaping photons from $\perp \rightarrow \parallel \parallel$ mode splitting, but only an attenuation cutoff in the $perp$ mode spectrum. Figure~\ref{fig:icspecpar} illustrates the effect of injecting photons at an angle (in this case $\theta_{\rm kB,0} = 0.57^\circ = 0.01$ radians) to the local magnetic field direction, toward the magnetic dipole axis. The high-energy cutoff decreases, compared to the case of injection parallel to the field, only in the $\perp$ mode spectrum and not at all in the $\parallel$ mode spectrum. This behavior is due to the existence of a threshold for pair production, but not for splitting and can be seen from Figs.~\ref{fig:icescape}a and~\ref{fig:icescape}b. For field strengths well above $B' = 0.1$, where photons pair produce at threshold, the pair escape energy is much less sensitive to increases in $\theta_{\rm kB,0}$ than is the splitting escape energy. The partial cascade spectra therefore become more highly polarized at small colatitudes when $\theta_{\rm kB,0}$ is increased. \placefigure{fig:icspecpar} This effect of strong polarization, both in the energy of the spectral cutoffs and the spectral shape just below the cutoffs, all but disappears when photon splitting is omitted from the calculation, thereby defining a characteristic signature of the action of \teq{\gamma\to\gamma\gamma}. Pair production has much less distinctive polarization features. For example, from equations~(\ref{eq:tpppar}) and~(\ref{eq:tppperp}), the ratio of the cutoff energies at pair creation threshold between the polarization states is \teq{(1+\sqrt{1+2B'})/2}. For surface fields of \teq{B'=0.7}, threshold is crossed during photon propagation in regions with much lower fields, typically \teq{B'\sim 0.1}, so that the spectral cutoff (or escape energy) differs only by about 5\% between polarizations; such a difference would be virtually invisible in the emission spectra. Clearly then, splitting is primarily responsible for polarization features shown. \subsection{Full Splitting Cascade} We now present model cascade spectra for the case where all three photon splitting modes allowed by CP invariance, $\perp \rightarrow \parallel \parallel$, $\perp \rightarrow \perp\perp$ and $\parallel \rightarrow \perp\parallel$, are operating, and multiple generations of splitting can occur. These cascades also allow for pair production by photons of either mode. As noted above, for the field strength of $B_0' = 0.7$ used in the spectral models for PSR1509-58, pair production occurs in less than $10\%$ of conversions. Figure~\ref{fig:specpol} shows full splitting cascade spectra in each final polarization mode for 2 million photons injected parallel to the local magnetic field (in curved spacetime) at different magnetic colatitudes. Each cascade spectrum shows a cutoff at roughly the splitting escape energy for that colatitude (compare to Fig.~\ref{fig:escape}), and a bump below the cutoff from the escaping cascade photons. The size of the bump is a function of the number of photons attenuated above the cutoff, which is dependent on the ratio of the maximum input energy, $\varepsilon_{max}$, and the escape energy. For these models, the size of the cascade bump grows with increasing $\theta$ because $\varepsilon_{max}$ is held constant while the escape energy is decreasing. The number of splitting generations ranges from 12 when $\theta = 30^\circ$ to 3 when $\theta = 2^\circ$. The size of the cascade bump at a particular $\theta$ could of course be larger or smaller if $\varepsilon_{max}$ were increased or decreased, but the positions of the cutoffs would not vary. The spectrum of the bump is polarized, with a well-defined zero in polarization that is a characteristic signature of the splitting cascade (see Baring 1995). Note that the polarization modes have reversed their flux dominance in the cascade bump compared to the partial splitting cascade case. \placefigure{fig:specpol} Although we have injected unpolarized photons in these calculations for simplicity, the relative flux (i.e. spectra integrated over energies) of the two polarization modes generally has a complicated dependence on the branching ratios for splitting defined by equation~(\ref{eq:splitpolrate}), due to the cascading process and the non-uniformity of the field. Notwithstanding, the polarization at a given energy does not exceed a limiting value of 3/7 (Baring 1991). The cascade spectra for injection of polarized photons resemble the spectra in Fig.~\ref{fig:specpol}, though deviations from Fig.~\ref{fig:specpol} are not exactly proportional to the degree of polarization of the injection spectrum due to the inherent complexity of the interplay of polarization states in the cascade. \placefigure{fig:icspecpol} As shown in Figure~\ref{fig:icspecpol}, injecting photons at an angle to the local field again has a much larger effect at small colatitudes (i.e. for $\theta \lesssim 100\theta_{\rm kB,0}$). The high-energy cutoffs in both modes now decrease in energy compared to the case of injection parallel to the field, and the sizes of the cascade bumps are larger, both being consequences of decrease in escape energy (see Fig.~\ref{fig:icescape}a). This effect is larger at smaller colatitudes. \section{PHOTON SPLITTING CASCADE MODELS FOR PSR1509-58} The multiwavelength spectrum of PSR1509-58, compiled from radio to TeV energies (Thompson 1996), shows that the peak in the power output from this pulsar, as is the case for most other $\gamma$-ray pulsars, falls in the $\gamma$-ray band. Figures~\ref{fig:datpar}--\ref{fig:daticunpol} show the high energy portion of this spectrum, near the cutoff, which we compare with our model spectra at different emission colatitudes. No formal procedure for fitting the data with the model was followed, since a simple visual comparison demonstrating the spectral cutoff is sufficient for the scientific goals of this paper. The $\epsilon^2\,F(\epsilon)$ format plots equal energy per logarithmic decade and clearly demonstrates the need for a cutoff or sharp turnover somewhere between the highest OSSE detected point at 3 MeV and the lowest EGRET upper limit at 30 MeV. Although there appears to be a discontinuity between the GINGA data points below 100 keV and the OSSE data points, it is common for separate fits of data from two different detectors to produce disparate results, even in the same energy range. Furthermore, the $\epsilon^2\,F(\epsilon)$ format tends to magnify the differences. The difference in spectral index of the Ginga and OSSE fits probably indicates a true break in the power-law spectrum around 100 keV. We have taken the OSSE index for the input spectrum for our cascade simulation since it most accurately measures the observed spectrum at the energies of importance for the model. The offset between the Ginga and OSSE data (or their different spectral indices) does not impact the conclusions of this paper, since the cascade formation is determined by the photon population in the upper end of the OSSE range. Note that while EGRET has obtained upper limits to the pulsed emission above around 30 MeV, there are earlier reports of a marginal detection by COS-B (e.g.. Hartmann et al. 1993), with data points lying above the EGRET limits. This apparent discrepancy remains to be resolved, and we opt here to consider only the later and superior EGRET observations. The Comptel point and limits in Figs.~\ref{fig:datpar}--\ref{fig:daticunpol} are a preliminary analysis of data from VP23 (Hermsen et al. 1996), showing pulsed flux at 0.75 - 1 MeV and upper limits for the pulsed interval (50\%) of the light curve. The cutoffs in the model photon splitting cascade spectra in Figures~\ref{fig:datpar}--\ref{fig:daticunpol} do in fact fall in the energy range 3--30 MeV for colatitudes less than around $30^\circ$. At colatitudes greater than $\sim 30^\circ$ the cutoffs are lower and are in severe conflict with the OSSE data points. The standard polar cap half-angle in flat spacetime, $\sin\theta = (\Omega R/c)^{1/2}$, for PSR1509-58 is $2.14^\circ$. Although curved spacetime corrections to the magnetic dipole field tend to very slightly decrease the polar cap size (GH94), the polar cap may be larger than the standard size due to distortion of the field near the light cylinder by plasma loading (Michel 1991). The results presented here assume, for simplicity, a single colatitude of emission for each, i.e. a polar rim rather than an extended cap. It is easy to envisage that a range of polar cap emission locations will produce a convolution of the spectra presented here, thereby generating a spectral turnover corresponding to the maximum colatitude of the cap, with steeper emission extending up to a cutoff defined by the minimum colatitude. The EGRET upper limits cannot really discern between a sharp cutoff or a more modest turnover above the Comptel energy range and so cast little light on the emission as a function of colatitude when \teq{\theta\lesssim 2^\circ}. \placefigure{fig:datpar} \placefigure{fig:daticpar} The partial splitting cascade spectra, shown in Figs.~\ref{fig:datpar} and~\ref{fig:daticpar}, exhibit only modest cascade bumps just below the cutoff. The limits on colatitude of the model spectra are essentially determined by the cutoff energy and are restricted by the lowest EGRET upper limit to $2^\circ \lesssim \theta \lesssim 25^\circ$ in the $\theta_{\rm kB,0} = 0$ case, and $\theta \lesssim 25^\circ$ in the $\theta_{\rm kB,0} = 0.57^\circ$ case, where no lower limit to the colatitude is imposed by the observations (see below). The model spectra for $\theta = 2^\circ$ and $5^\circ$ are only marginally consistent with the upper limits. The final revision of the Comptel data for PSR1509-58 (Bennett et al., in preparation) may require raising the lower bounds to the colatitude of emission obtained in this model/data comparison. The cutoff energies of these polarization-averaged spectra are somewhat larger than the cutoff energies of the full cascade spectra (see Figs.~\ref{fig:datunpol} and~\ref{fig:daticunpol}) because the $\parallel$ mode escape energies are determined solely by pair production, whose escape energies generally exceed those of splitting at this field strength (see Fig.~\ref{fig:escape}). This is especially pronounced in the $\theta_{\rm kB,0}=0.57^\circ$ case, due to the fact that the pair production escape energy is insensitive to the photon emission angle for $B \gg 0.1$, as is illustrated in Fig.~\ref{fig:icescape}b. The full cascade spectra, shown in Figs.~\ref{fig:datunpol} and~\ref{fig:daticunpol}, have distinctive bumps below the cutoff due to the redistribution of photon energies via splitting. The size of the cascade bump further limits the magnetic colatitudes to $\theta \lesssim 5^\circ$ to avoid conflict with the Comptel upper limits. The lowest EGRET upper limit restricts the colatitudes to $5^\circ \gtrsim \theta \gtrsim 2^\circ$ in the case of emission parallel to the field (Fig.~\ref{fig:datunpol}). In the case of emission at angle $\theta_{\rm kB,0} = 0.57^\circ$ (Fig.~\ref{fig:daticunpol}), the cutoff energy in the cascade spectra saturates at small $\theta$ at an energy of 25 MeV (see Fig.~\ref{fig:icescape}a). Consequently there is no low-energy limit to $\theta$ in this case. For larger values of $\theta_{\rm kB,0}$, the spectral cutoffs would saturate at larger values of $\theta$ and at lower energies. We can estimate the dependence of this saturation escape energy, $\varepsilon_{\rm esc}^{\rm sat}$, on $\theta_{\rm kB,0}$ and $B'$ using the expression for the splitting attenuation coefficient in equation~(\ref{eq:splitotrate}). Assuming that $1/T_{\rm sp} \simeq R$ approximately gives the escape energy: \begin{equation} \varepsilon_{\rm esc}^{\rm sat} \simeq 0.077\,(B'\sin\theta_{\rm kB,0})^{-6/5} \;\; ,\quad B' \lesssim 0.3. \label{eq:ergs} \end{equation} This formula quite accurately reproduces the escape energies in Figs.~\ref{fig:datunpol} and~\ref{fig:daticunpol} since they are only weakly dependent on \teq{R}, specifically \teq{ \varepsilon_{\rm esc}^{\rm sat}\propto R^{-1/5}}. When $\varepsilon_{\rm esc}^{\rm sat} \le 7.8$ (i.e. 4 MeV), cascade spectra at all colatitudes cutoff below the lowest possible observed cutoff energy for the PSR1509-58 spectrum. From equation~(\ref{eq:ergs}), this occurs, for surface emission, at $\theta_{\rm kB,0} \gtrsim 0.03$ for $B_0' = 0.7$. Therefore, splitting cascade spectra from photons emitted at larger angles to the field will not be compatible with the spectrum of PSR1509-58. For emission at some distance above the surface, the limit on $\theta_{\rm kB,0}$ would be higher since it depends inversely on local field strength. \placefigure{fig:datunpol} \placefigure{fig:daticunpol} All the model spectra in Figures~\ref{fig:datpar}--\ref{fig:daticunpol} assume emission at the neutron star surface. Emission above the surface would produce higher cutoff energies at a given colatitude, due to the decrease in the dipole field strength with $r$. The upper limits on colatitude stated above would therefore be less restrictive for non-surface emission. Furthermore, when the field strength at the emission point is $B \sim 0.3 B_{\rm cr}$ (at height $30\%$ of the neutron star radius) the splitting and pair production escape energies are comparable, reducing the size of the splitting cascade bumps in all cases. At higher altitudes above the surface, pair production dominates the photon attenuation and conventional pair cascades (e.g. Daugherty \& Harding 1996) would operate. Synchrotron radiation from the pairs would then result in a significantly softer emergent spectrum than the input power-law above the cyclotron energy (\teq{\sqrt{1+2B'}-1\approx 280}keV at the stellar surface, lower at greater radii). Consequently, in order to match the observations, the input power-law would have to be harder, and because of the remoteness of the emission point from the stellar surface, the colatitude \teq{\theta} of emission would have to be increased substantially. \section{DISCUSSION} The results of this paper demonstrate that magnetic photon splitting can have a significant effect on $\gamma$-ray emission from the higher field ($B_0 \gtrsim 10^{13}$ G) pulsars. It can attenuate the $\gamma$-ray spectrum at lower energies than magnetic pair production and will do so without the creation of electron-positron pairs. We have found that in low fields ($B_0 \lesssim 0.3B_{\rm cr}$) and $\theta_{\rm kB,0} = 0$ initially, photon splitting attenuation lengths are never shorter than those for pair production. In high fields ($ B_0 \gtrsim 0.3B_{\rm cr}$), photon splitting lengths fall below those for pair production below a certain energy which depends on the colatitude $\theta$. Photon splitting escape energies fall below pair production escape energies for $B_0 \gtrsim 0.5\,B_{\rm cr}$, so that splitting may produce an observable signature for $\gamma$-ray pulsars having strong magnetic fields: high energy spectral cutoffs that are quite polarization-dependent. While pair creation alone will also generate such cutoffs, their dependence on photon polarization is far diminished from when splitting is active. We have modeled the shape of such spectral cutoffs through simulation of photon splitting cascades near the neutron star surface for the case of PSR1509-58. Two types of cascades result from different assumptions about the selection rules governing the photon splitting modes: the ``full splitting cascades" occur when three modes limited only by CP selection rules operate and the ``partial splitting cascades" occur when only one mode permitted by kinematic selection rules operates. In the full cascades, splitting dominates the attenuation while in the partial cascades, pair production ultimately limits the rate at which photon energy degrades. However, the partial cascades show a distinct polarization signature due to the different escape energies for splitting and for pair production. The resulting PSR1509-58 model spectral cutoffs due to splitting and pair production fall in the required range for virtually all colatitudes $\lesssim 25^\circ$. However, the shape of the spectrum of full splitting cascades, due to the large reprocessing bump, is compatible with the data only for a very small range of colatitudes, $\theta \lesssim 5^\circ$. From these results we conclude that, although photon splitting is capable of producing spectral cutoffs well below EGRET energies regardless of which selection rules govern the splitting modes, the partial splitting cascades have a much larger range of phase space in which to operate. Attenuation through magnetic pair production and photon splitting near the polar cap will produce $\gamma$-ray spectral cutoffs that should be roughly a function of surface magnetic field strength, although other parameters such as polar cap size will come into play. Thus the $\gamma$-ray pulsar PSR0656+14, having the second highest surface field of $9.3 \times 10^{12}$ G, should have a cutoff energy between that of PSR1509-58 and the other $\gamma$-ray pulsars. In fact the unusually large spectral index of 2.8 measured by EGRET (Ramanamurthy et al. 1996) may be a pair production/photon splitting cutoff. It is thus possible to understand why PSR1509-58, with the highest magnetic field of all the $\gamma$-ray pulsars, has by far the lowest spectral cutoff energy and is the only $\gamma$-ray pulsar not detected by EGRET. In the case of Vela (Kanbach et al. 1994), Geminga (Meyer-Hasselwander et al. 1994) and 1055-52 (Fierro et al. 1993), the spectral cutoffs observed by EGRET at a few GeV are consistent with one-photon pair production cascades (Daugherty and Harding 1982, 1996). Although the escape energies at the neutron star surface for the spin-down fields of these pulsars ($B_0 \sim 2 - 6 \times 10^{12}$ G) is below 1 GeV (see Fig.~\ref{fig:icescape}a), curvature radiation from primary electrons at one to two stellar radii above the surface will have pair production escape energies of several GeV. However, when the surface field exceeds $\sim 10^{13}$ G, photon splitting becomes the dominant attenuation mechanism in the electromagnetic cascades. In addition, the primary electrons may lose energy to resonant Compton scattering of thermal X-rays from the neutron star surface (Sturner 1995), rather than to curvature radiation, limiting their acceleration to much lower energies, typically $\gamma \sim 100$. The resulting upscattered $\gamma$-ray spectrum is radiated much closer to the surface and will be cut off by photon splitting well below the EGRET energy range. It is important to emphasize that pair creation acting alone suffices to inhibit GeV emission in pulsars with spin-down fields as high as PSR1509-58, and splitting significantly enhances the attenuation and pushes spectral cutoffs to lower energies. If resonant Compton scattering losses limit the polar cap particle acceleration energies to $\gamma \ll 10^6$ when $B \gtrsim 10^{13}$ G, then the primary particles will radiate $\gamma$-rays via the cyclotron upscattering process or CUSP (Dermer 1990). CUSP radiation would then provide the seed photons for the splitting cascade. The $\gamma$-ray spectrum for this process for power-law and monoenergetic electrons scattering thermal blackbody X-ray photons above the neutron star surface (Daugherty \& Harding 1989) is a power-law with maximum energy $\varepsilon_{max} \simeq \gamma_c B' = 2 \times 10^3\,B'^2/T_6$ (Dermer 1990), where $\gamma_c$ is the energy above which the electrons scatter resonantly and $T_6 \equiv T/10^6$ K is the thermal X-ray temperature. In the case of PSR1509-58 with $B' = 0.7$, $\varepsilon_{max} \simeq 10^3/T_6$. Since the thermal surface emission component is not observed due to the strong non-thermal spectrum seen at X-ray energies (Kawai 1993), $T_6$ is not known. However, PSR1509-58 is young ($\sim 1000$ yr) and probably has $T_6 \sim 1 - 3$. We would then expect $\varepsilon_{max} \simeq 300 - 10^3$, compatible with our choice of $\varepsilon_{max} = 100$ for the splitting cascade models. A dozen or so other radio pulsars have spin-down magnetic fields above $10^{13}$ G. These pulsars would, like PSR1509-58, have photon splitting dominated cascades rather than pair cascades, producing lower yields of electron-positron pairs. It is possible that neutron stars with extremely high magnetic fields, where splitting is dominant at altitudes up to several stellar radii, do not produce sufficient pairs for coherent radio emission, an intriguing possibility. If such neutron stars exist, they would constitute a new class of radio quiet, low-energy $\gamma$-ray pulsars. \acknowledgements We thank Dieter Hartmann and David Thompson for reading the manuscript and for providing helpful comments, and Wim Hermsen, Alberto Carrami\~nana and Kevin Bennett for providing preliminary Comptel data for PSR1509-58. This work was supported through Compton Gamma-Ray Observatory Guest Investigator Phase 5 and NASA Astrophysics Theory Grants. MGB thanks the Institute for Theoretical Physics at the University of California, Santa Barbara for support (under NSF grant PHY94-07194) during part of the period in which work for this paper was completed. \clearpage
proofpile-arXiv_065-499
{ "file_path": "/home/ubuntu/dolma-v1_7/algebraic-stack-train-0000.json.gz" }